You are on page 1of 183

An Introduction to

Safety Grounding
An Introduction to
Safety Grounding

Asser A. Zaky
First edition published 2022
by CRC Press
6000 Broken Sound Parkway NW, Suite 300, Boca Raton, FL 33487-2742

and by CRC Press


2 Park Square, Milton Park, Abingdon, Oxon, OX14 4RN

© 2022 Taylor & Francis Group, LLC

CRC Press is an imprint of Taylor & Francis Group,LLC

The right of Asser A. Zaky to be identified as author of this work has been asserted by him in
accordance with sections 77 and 78 of the Copyright, Designs and Patents Act 1988.

Reasonable efforts have been made to publish reliable data and information, but the author and
publisher cannot assume responsibility for the validity of all materials or the consequences of
their use. The authors and publishers have attempted to trace the copyright holders of all material
reproduced in this publication and apologize to copyright holders if permission to publish in this
form has not been obtained. If any copyright material has not been acknowledged please write and
let us know so we may rectify in any future reprint.

Except as permitted under U.S. Copyright Law, no part of this book may be reprinted, reproduced,
transmitted, or utilized in any form by any electronic, mechanical, or other means, now known or
hereafter invented, including photocopying, microfilming, and recording, or in any information
storage or retrieval system, without written permission from the publishers.

For permission to photocopy or use material electronically from this work, access www.copyright.
com or contact the Copyright Clearance Center, Inc. (CCC), 222 Rosewood Drive, Danvers, MA
01923, 978-750-8400. For works that are not available on CCC please contact mpkbookspermissions@
tandf.co.uk

Trademark notice: Product or corporate names may be trademarks or registered trademarks and are
used only for identification and explanation without intent to infringe.

Library of Congress Cataloging‑in‑Publication Data


Names: Zaky, Asser A. (Asser Aly), 1932- author.
Title: An introduction to safety grounding / Asser A. Zaky.
Description: First edition. | Boca Raton, FL: CRC Press, 2021. | Includes bibliographical references
and index. | Summary: “Protective or safety grounding is of vital importance for the protection of
individuals from electric shock. The objective of this book is to give the reader a better understanding
of safety grounding, why it is needed, where it is needed, and what are the requirements which must
be met in order to have an effective grounding system”—Provided by publisher.
Identifiers: LCCN 2021004135 (print) | LCCN 2021004136 (ebook) |
ISBN 9780367758714 (hbk) | ISBN 9780367759278 (pbk) |
ISBN 9781003164630 (ebk)
Subjects: LCSH: Electric currents—Grounding. |
Electric wiring—Insurance requirements.
Classification: LCC TK3227 .Z35 2021 (print) | LCC TK3227 (ebook) | DDC 621.31/7—dc23
LC record available at https://lccn.loc.gov/2021004135
LC ebook record available at https://lccn.loc.gov/2021004136

ISBN: 978-0-367-75871-4 (hbk)


ISBN: 978-0-367-75927-8 (pbk)
ISBN: 978-1-003-16463-0 (ebk)

Typeset in Times
by codeMantra
Contents
Preface.......................................................................................................................ix

Chapter 1 Effect of Current on the Human Body .................................................1


1.1 Introduction ...............................................................................1
1.2 Effect of Electric Shock on Human Beings...............................3
1.3 Effect of Current Duration ........................................................4
1.4 The Electric Resistance of the Human Body ............................5

Chapter 2 Resistance to Ground ...........................................................................7


2.1 Resistance to Ground of a Hemisphere .....................................7
2.2 Measurement of Electrode Resistance to Ground ................... 11
2.3 Soil Resistivity ......................................................................... 16
2.4 Measurement of Soil Resistivity.............................................. 18
2.4.1 Simple Box Method .................................................... 18
2.4.2 Wenner Method .......................................................... 19
2.4.3 Electromagnetic Induction Method ............................20
2.5 Computer Programs................................................................. 22

Chapter 3 Earthing Electrodes ............................................................................ 23


3.1 Single Driven Rod ................................................................... 23
3.2 Grounding Using Multiple Rods .............................................25
3.3 Grounding Using Horizontally Buried Wires ......................... 30
3.4 Grounding Using Buried Plates............................................... 32
3.5 Water Pipes or Steel Reinforcement as
Grounding Electrodes ........................................................33
3.6 The Resistance Area ................................................................ 33

Chapter 4 Step, Touch, and Transfer Voltages .................................................... 35


4.1 Step Voltage ............................................................................. 35
4.2 Touch Voltage .......................................................................... 36
4.3 Transfer Touch Voltage ............................................................ 39
4.4 Grounding of Power Towers ....................................................40
4.4.1 Effect of Overhead Ground Wire on Tower
Footing Resistance ..................................................... 40

Chapter 5 Grounding Systems ............................................................................ 45


5.1 Magnitude of Resistance to Ground ........................................ 45
5.2 Elements of the Grounding System ......................................... 45

v
vi Contents

5.2.1 The Area of Land ....................................................... 47


5.2.1.1 Use of Chemical Salts ................................. 47
5.2.1.2 Increasing the Soil’s Ability to
Retain Water ............................................... 49
5.2.2 Ground Electrodes...................................................... 50
5.2.2.1 Driven Rods ................................................ 50
5.2.2.2 Concrete-Encased Electrodes ..................... 52
5.2.2.3 Water and Gas Pipes ................................... 53
5.2.2.4 Structural and Reinforcing Steel ................ 54
5.2.2.5 Equipotential Bonding ................................ 54
5.2.3 Connections in Grounding Systems ........................... 54
5.2.4 Grounding Conductors ............................................... 56
5.2.4.1 Size of Grounding Conductors ................... 57
5.2.4.2 Protective Sleeve ......................................... 59
5.2.5 Ground Connection Provisions ..................................60
5.2.6 Equipment Grounding Conductors ............................60
5.2.6.1 Conductor Size according to the
Adiabatic Equation ......................................61
5.2.6.2 Conductor Size According to USA
Specifications (NFPA-70) ............................67
5.2.6.3 Metal Cable Trays as Grounding
Conductors .................................................. 68
5.2.6.4 Flexible Metal Conduits.............................. 71
5.2.6.5 Cable Sheath and Armor ............................ 72
5.2.6.6 Jumpers ....................................................... 72
5.3 Metallic Corrosion................................................................... 72
5.4 General Rules for the Grounding of Branch Substations ........ 74
5.5 Protective Grounding of Consumer Installations:
Types of Grounding Systems....................................................79
5.5.1 The TN-C System ......................................................80
5.5.2 The TT System ........................................................... 82
5.5.3 The TN-S System .......................................................84
5.5.4 Protective Multiple Earthing ...................................... 87
5.5.5 The IT System ............................................................ 89
5.6 Protection by Isolation .............................................................90
5.7 Grounding of Computers and Data Processing Equipment ..........91

Chapter 6 Substation Grounding Systems .......................................................... 93


6.1 Introduction ............................................................................. 93
6.2 Design Steps ............................................................................94
6.2.1 Ground Mat Resistance ..............................................94
6.2.2 Calculation of Maximum GPR...................................97
6.2.3 Determination of Maximum Touch and
Step Voltages .............................................................. 98
Contents vii

6.2.4 Safety Requirements ................................................ 102


6.3 Design of Ground Mat using Computer Program ................. 104

Chapter 7 Static Electrification ......................................................................... 111


7.1 Introduction ........................................................................... 111
7.2 Conditions Necessary for Ignition ......................................... 111
7.2.1 Explosion Limits ...................................................... 112
7.2.2 Oxygen ..................................................................... 114
7.2.3 Ignition Sources........................................................ 114
7.2.4 Minimum Ignition Energy ....................................... 115
7.3 Generation of Electrostatic Charges ...................................... 115
7.3.1 Contact Electrification between Solid Surfaces ....... 117
7.3.2 Static Electrification in Liquids and Gases .............. 119
7.4 Ways for Reducing the Formation of Surface Charges ......... 120
7.5 Electrostatic Induction ........................................................... 122
7.6 Types of Electric Discharges ................................................. 122
7.6.1 The Spark or Arc Discharge..................................... 123
7.6.2 Brush Discharges and Propagating
Brush Discharges.......................................................127
7.6.3 Corona Discharges ................................................... 128
7.7 Methods of Controlling Electrostatic Charges ...................... 129
7.7.1 Grounding and Bonding ........................................... 129
7.7.2 Floors and Clothing for Static Control ..................... 130
7.7.2.1 Static Conductive Floors ........................... 131
7.7.2.2 Static Dissipative Floors ........................... 131
7.7.3 Control of Humidity ................................................. 133
7.7.4 Ionization .................................................................. 134

Chapter 8 Protection against Lightning ............................................................ 137


8.1 Nature of Lightning ............................................................... 137
8.2 Why Protect against Lightning.............................................. 140
8.3 Lightning Protection System ................................................. 144
8.3.1 The Air Terminals .................................................... 145
8.3.2 Mesh Air Termination Networks ............................. 146
8.3.3 Down Conductors ..................................................... 147
8.3.4 Types of Conductors ................................................. 149
8.3.5 Grounding ................................................................ 150
8.4 Inductance, Side Flashing, and Separation Distance ............ 150
8.5 Zones of Protection ............................................................... 153
8.6 Tanks and Stacks ................................................................... 157
8.7 Protection of Transmission Lines by Aerial Ground Wires .. 157
8.7.1 Secondary Field of Ground Conductor .................... 158
8.7.2 Height of Ground Conductor .................................... 161
8.8 Protection against Surges ...................................................... 161
viii Contents

Appendix A: Wire Sizes....................................................................................... 163


Bibliography.......................................................................................................... 167
Index....................................................................................................................... 169
Preface
It has often been said that grounding is part science and part art. To a large extent this
is true if we consider the art part to include the extensive experience accumulated over
centuries, going back to Benjamin Franklin’s experiments with kites and lightning
and his invention of the lightning rod, which gave birth to the concept of grounding
as a safety measure.
As with many engineering inventions, theory followed practice and the com-
bined outcome of both theory and practical experience have been embodied in a
number of national and international standards and codes of practice on safety
grounding, which are revised from time to time to take into account the develop-
ment of new materials, improvements in engineering practice, and new research
findings.
Ever since grounding was used as a means of protection against lightning, its
importance as a safety measure in electrical engineering applications, from genera-
tion to utilization, has been universally recognized and is today an obligatory safety
requirement for all electrical installations and all nonbattery operated electrical
equipment. Unfortunately the implementation of this requirement is more often than
not overlooked in many developing countries. The reason for this is a mixture of
ignorance and a laxity in the enforcement of safety measures. The majority of engi-
neering curricula does not explicitly address safety grounding but emphasizes system
grounding. It is not surprising therefore that engineers do not always distinguish
between these two types of grounding. This lack of awareness applies also to techni-
cians and wiring electricians. There certainly is a need for a wider understanding of
safety grounding and its importance as a life-saving measure.
During years of teaching electrical power and high voltage courses I have always
made it a point to include safety grounding as an integral part of the course. The pres-
ent text is the outcome of these lecture notes enlarged and updated as far as possible
to conform to present-day standards and practice.
The choice of material for a balanced textbook on grounding is not easy.
A  comprehensive treatment of each individual topic requires a book by itself; for
instance, Tagg’s classic book Earth Resistances is almost exclusively devoted to earth
electrodes, Golde’s Lightning Protection to the protection against lightning, Loeb’s
Static Electrification to the triboelectric effect to quote but a few. In the present text
the topics and depth of treatment have been chosen with the prime objective of giving
the reader a simple theoretical background to each topic and a “feel” for the design
of practical grounding systems which is intimately and inexorably bound to existing
standards, both national and international. Hence reference to such standards and
their requirements is frequently made throughout the book.
The book addresses students and professionals interested to learn more about
safety grounding other than that it is a piece of wire with which equipment is
connected to ground, especially since the very numerous standards on every aspect
of the subject are not readily accessible.

ix
x Preface

The book has eight chapters:


Chapter 1 gives the physiological effects which the magnitude of current and its
duration has on the human body.
Chapter 2 deals with the resistance of an electrode to ground and the resistivity of
the soil. Methods of measuring this resistance as well as the resistivity of the soil are
given. The resistance area of a ground electrode is defined.
Chapter 3 deals with the different types of ground electrodes and the effect of
their geometry and numbers on the resistance to ground.
Chapter 4 covers the concepts of step, touch, and transfer voltages and their safe values.
Chapter 5 is the core of the book and presents in some detail the components
of a ground system, methods of improving soil resistivity, the types of welds and
joints,the criteria for determining conductor cross-sections, galvanic corrosion, and a
survey of the different grounding practices used at substations and the different types
of grounding systems used for the protection of consumers.
Chapter 6 gives a concise treatment of the use and design of substation ground
mats such that step and touch voltages remain within the safe limits and end with
examples using a computer program for the design of ground mats.
Chapter 7 deals in some detail with static electrification and the types of electrostatic
discharges (ESD) for which they are responsible, especially since such discharges
have today acquired considerable importance in the electronics manufacturing
industries. The chapter includes the various methods used to minimize or prevent the
occurrence of such discharges.
Chapter 8 introduces the reader to the subject of lightning and lightning protection.
Lightning of course is as old as the earth itself and to this day many aspects remain
unsolved with research still continuing. However, the protective measures used so
far have proved to be quite effective and are continuously being improved upon as
evidenced by the recent changes brought about in almost all relevant standards.
It should be pointed out that the use of “ready-made” computer programs for
the design of a ground system may be labor saving but it is the author’s opinion
that a sound knowledge of the premises on which such programs are designed is of
fundamental importance for their proper use.
It is hoped that the coverage provided in the present monograph will help students
and engineers alike to better understand protective grounding and the standards
involved and appreciate its importance and so champion its implementation in their
field of work and thereby be blessed for a life-saving act.
In preparing this text recourse has been had to many standards and references, and
wherever necessary, tables and figures have been reproduced with due acknowledgment
of their source. In particular, the author would like to thank W.J.Furse & Co for
permission to use a number of their figures and the NFPA for permission to use
their simple lightning risk assessment procedure. The author thanks the International
Electrotechnical Commission (IEC) for permission to reproduce information from its
International Standards.1 Thanks are also due to the publishing staff of T&F for their
assistance and guidance throughout the preparation of this book.

1 All such extracts are copyright of IEC, Geneva, Switzerland. All rights reserved. Further information
on the IEC is available from www.iec.ch. IEC has no responsibility for the placement and context in
which the extracts and contents are reproduced by the author, nor is IEC in any way responsible for the
other content or accuracy therein.
1 Effect of Current on
the Human Body

1.1 INTRODUCTION
There are essentially two types of grounding:1

1. System grounding
2. Protective or equipment grounding

In the first type of grounding the star points of the equipment may be solidly grounded
or grounded through a resistance or an inductance (Petersen coil) according to the
operating requirements of the network. These requirements depend on several factors
such as the maximum permissible stress on the insulation, the magnitude of the short
circuit current, and the overall protective characteristics of the network.
The second type of grounding, which is the subject of this book, has two objectives:

(ii) To protect buildings and installations against fire and lightning.

In order to protect people (operators, maintenance and repair technicians, and the
public at large) against electric shock if they come in contact with metal parts which
normally are not live and do not carry any current, it must be ensured that under
fault conditions the potential of such parts does not rise to a value which would be
considered dangerous to persons or give rise to leakage currents which, even if very
small, can with time raise the temperature of the material through which it flows to
a value sufficient to initiate a fire if the material is readily inflammable. Protection
is provided by deliberate grounding of all metal structures, motor, generator and
transformer frames, metal enclosures of all tools and control equipment, connection
boxes, cable trays, and all other metal bodies which contain or are adjacent to electric
circuits and which are within reach of any person.
Figure 1.1 shows the equivalent circuit of an electrical equipment connected to a
supply source of voltage V. The resistances shown are as follows:
R1: resistance of the insulation between live parts and equipment case,
Rg: resistance between casing and ground, and
Rb: resistance of a person’s body to ground.

1 The terms grounding and earthing are used interchangeably.

1
2 An Introduction to Safety Grounding

L R1
Vt
V ZL leakage
N current Rb

Rg

FIGURE 1.1 Isolated equipment.

The voltage Vt which appears on the case is

V Rg
Vt = (1.1)
R1 + Rg

When a short circuit occurs between the live conductor and the case (R1 = 0), Vt
becomes equal to V and the current which flows through the body of a person who
touches the frame is

I b = V /Rb .

If the case were connected to a perfect ground such that Rg = 0 then the case would
always be at zero potential. However, in practice the resistance to ground is never
zero so that under fault conditions the case voltage is V and remains at that value until
the protective device (fuse or circuit breaker) disconnects the supply (Figure 1.2). To
make certain that the person who touches the case is not at risk the circuit must be
disconnected within a specified time (see Section 1.3); since the operating time of the
protective devices depends on the magnitude of the short circuit current, it is neces-
sary to ensure that the resistance between the case and ground is sufficiently small
to allow the passage to ground of a current whose value is sufficient to operate the

V
fuse or CB

ZL

FIGURE 1.2 Earthed equipment.


Effect of Current on the Human Body 3

protective device within the permitted time. To fulfill this requirement the resistance
to ground must not exceed a certain value usually specified by national or interna-
tional standards. This value varies between 1 and 25 Ω depending on the magnitude
of the short circuit current.
As for the protection of buildings and installations against lightning strikes, this is
accomplished by earthing systems especially designed for this purpose and its imple-
mentation is determined by how critical such a protection is. Protection of structures
against lightning is dealt with in Chapter 8.
In all of the national and international specifications protective grounding of all
equipment is an obligatory safety requirement irrespective of other considerations
such as electromagnetic compatibility, for example. It is therefore of primary impor-
tance that there be a close cooperation between the engineering consultant responsi-
ble for the design of supply and grounding system of any installation and the designer
of the electronic equipment to be installed in the building in order to choose the
optimum system which will meet both the safety requirements and the compatibility
requirements within the electromagnetic environment at the premises.

1.2 EFFECT OF ELECTRIC SHOCK ON HUMAN BEINGS


A person is subjected to an electric shock if he touches any live conductor or other
metallic body while at the same time touching another grounded body or standing
on a moist ground or is barefooted. Although the common belief is that the voltage is
the cause of the shock, the consequences and severity of the shock depend on many
factors, the most important of which is the magnitude of the current which flows
through the body, the path of this current, and its duration.
Numerous studies and observations have shown that the effect of a low-frequency
current (0–300 Hz) on the human body varies with the magnitude of the current which
flows through the chest area. Table 1.1 shows a summary of the effects produced by
both alternating and direct current magnitudes for durations of a few seconds. Note
that the current magnitude is in milliamperes.
The minimum current which can be felt by human beings is one milliampere.
Between 6 and 10 mA muscle control is lost so that a person cannot let go of any

TABLE 1.1
Effects of Current Magnitude on the Human Body
AC DC
mA, 50 Hz mA Effect
0.5–1 1–4 Threshold of feeling
1–10 4–15 Pain
10–30 15–80 Let-go threshold
30–50 80–160 Muscular paralysis
50–75 160–300 Difficulty in breathing
75–250 300–500 (Ventricular fibrillation) death
> 250 > 500 Cardiac arrest and serious burns
4 An Introduction to Safety Grounding

electrified body held in his hand. As the current magnitude increases breathing
becomes difficult and the muscles become paralyzed. These effects are not permanent
and disappear if the current is switched off within a few seconds. Even if breathing
stops the injured person can be saved from suffocation by artificial respiration.
However, if the current is between 75 and 250 mA the electric shock is fatal. The
reason for this is that within this current range the heart goes into a state know as
ventricular fibrillation in which the heart muscles no longer contract in synchronism.
This state is called cardiac arrest and it is more dangerous as it can only be reversed
by specialized equipment only available in hospitals. Currents above 250 mA lead to
cardiac arrest, cessation of breathing, and severe burns; however if the injured person
is given immediate treatment resuscitation is possible.
In the case of direct current, the current magnitudes producing the above effects
vary between two and four times the AC values as shown in Table 1.1.
At high frequencies the magnitude of the current required to produce the above
effects increases with increasing frequency due to the skin effect. For example the
threshold of feeling at 70 kHz is 100 mA and for frequencies higher than 100–200 kHz
the effect is limited to a sensation of heat or to the occurrence of superficial burns.

1.3 EFFECT OF CURRENT DURATION


In order to prevent an electric shock from causing death the magnitude of the current
must be less than that which causes ventricular fibrillation. As mentioned above the
magnitude of this current varies between 75 and 250 mA (the actual value depends on
the size of the body) if the duration of the current is a few seconds. Experiments and
statistics indicate that the shorter the duration of the current the higher the current
needed to cause fibrillation. Although there is complete agreement between inves-
tigators on the importance of current duration, there is no law or equation agreed
upon internationally for relating the magnitude of the current causing fibrillation to
the duration of that current. The most common relationship used is that arrived at
by Dalziel2 and modified more recently by others3 on the basis of numerous experi-
ments. This relationship is

Ib = k / t ampere (1.2)

where t is the duration of the shock current in seconds and k is a constant whose value
depends on the weight of the person:

k = 0.116 for 99.5% of persons weighing 50 kg


k = 0.157 for 99.5% of persons weighing 70 kg
If we assume that k = 0.116 then,
Ib= 116 mA t = 1 s
Ib= 367 mA t = 0.1 s

2 G.F. Dalziel, Dangerous electric currents, AIEE Trans.Vol.65, pp 579–585, and 1123–1124, 1946.
3 J.G. Sverak, W.K. Dick, T.H. Dodds and R.H. Heppe, Safe substation grounding – Part I, IEEE Trans.
PAS, pp 4281–4284, 1981.
Effect of Current on the Human Body 5

a b c1 c2 c3

Duration of current, ms

Body current, mA

FIGURE 1.3  Conventional time/current zones of effects of AC currents (15–100 Hz)


on ­persons for a current path corresponding to left hand to feet according to IEC 60479-1
(see Table 1.2).

Equation (1.2) is based on experiments in which the duration of the current varied
between 0.03 and 3 seconds and on the assumption that the resistance of the human
body is 1,000 Ω (see Section 1.4). This equation shows that higher shock currents are
allowed if high-speed protective devices are used to disconnect the current and limit
its duration. If the probability exists of prolonged shock duration without immediate
aid, then to prevent the injured person from choking due to cessation of breathing, it
is preferable to limit the maximum shock current to 25 mA. However, most designers
prefer to limit this current to the let-go value (9 mA for men and 6 mA for women).
Figure 1.3 gives the physiological effects of alternating current (15–100 Hz) dura-
tion on the human body according to the International Electrotechnical Commission
(IEC) standard,4 and a summary of the time zones is given in Table 1.2.

1.4 
THE ELECTRIC RESISTANCE OF THE HUMAN BODY
It is difficult to specify an exact value for the resistance of the human body since
its value is determined by several factors such as age, sex, weight, general physical
condition, extent of skin dryness, and the position of the body at the time of the acci-
dent. The total resistance of the body consists of two parts: the skin resistance and
the internal resistance.
The skin resistance varies between 100 Ω/cm2for wet skin and 3 × 105Ω/cm2 for
dry skin with values higher than that for people with coarse hands such as manual
workers.
The internal resistance of the body is in the range 400–600 Ω between
­extremities—hand to hand, hand to foot, or foot to foot.

4 IEC 60479-1 ed.1.0 “Copyright © 2018 IEC Geneva, Switzerland. www.iec.ch”


6 An Introduction to Safety Grounding

TABLE 1.2
Time/Current Zones for AC 15–100 Hz for Hand to Feet Pathway (Summary
of Zones in Figure 1.3)
Zones Boundaries Physiological Effects
AC-1 Up to 0.5 mA Perception possible but usually no “startled reaction”
curve a
AC-2 0.5 mA up to Perception and involuntary muscular contractions likely but usually
curve b no harmful electrical physiological effects
AC-3 Curve b and Strong involuntary muscular contractions. Difficulty inbreathing.
above Reversible disturbances of heart function.
Immobilization may occur. Effects increasing with current
magnitude. Usually no organic damage expected.
AC-4a Above Pathophysiological effects may occur such as cardiac arrest, breathing
curve c1 arrest, and burns or other cellular damage. Probability of ventricular
fibrillation increasing with current magnitude and time.
c1–c2 AC-4.1 Probability of ventricular fibrillation increasing up to 5%.
c2–c3 AC-4.2 Probability of ventricular fibrillation up to about 50%.
Beyond AC-4.3 Probability of ventricular fibrillation above 50%.
curve c3

a For durations of current flow below 200 ms, ventricular fibrillation is only initiated within the vulnerable
period if relevant thresholds are surpassed. With regard to ventricular fibrillation, this figure relates to
the effects of current which flows in the path left hand to feet. For other current paths, the heart current
factor has to be considered.

Although there is a large discrepancy in the value of the overall body resistance,
measurements and experience have shown that a suitable value for this resistance is
1,000 Ω between extremities (hand to hand, hand to both feet, or from one foot to
the other). This is the value adopted by the majority of standards and it is the one we
shall use throughout this book.
2 Resistance to Ground

2.1 RESISTANCE TO GROUND OF A HEMISPHERE


The hemisphere is the simplest geometrical shape of an earth electrode whose
resistance to ground can be easily calculated. If we assume that the region around
the hemisphere is divided into concentric hemispherical shells of equal thickness
dx (Figure 2.1) and that the earth resistivity is uniform, it is apparent that the shell
nearest to the electrode has the greatest resistance since it has the smallest area nor-
mal to the current flow. Each successive shell has a larger area and hence a smaller
resistance. The total resistance between the electrode and ground can be determined
as follows. Let a be the radius of the hemisphere, I the current flowing to ground, and
ρ (Ω·m) the earth resistivity. The resistance of a shell of radius x and thickness dx is

dR = ρ dx /2πx 2

and the resistance between the surface of the electrode and a point at a radial distance
r from its center is
r
ρ 1 1
R =
∫ dR = 2π  a − r 
a
(2.1)

If r becomes infinite the absolute resistance to ground of the hemisphere is

ρ
Rg = ohms (2.2)
2π a

r x dx
a

FIGURE 2.1 Hemispherical ground electrode.

7
8 An Introduction to Safety Grounding

Figure 2.2 shows the variation of the resistance to ground as given by Eq. (2.1). From
this equation it is apparent that 90% of the absolute resistance of the electrode to
ground lies in the region around the electrode whose radius is ten times the radius of
the electrode itself.
The potential difference between the electrode and a point at a distance r from its
center is given by

Iρ Iρ
Var = IR = − (2.3)
2π a 2π r

Since the absolute potential at any point is the potential at that point with respect to a
point at infinity (zero potential), the absolute potential of the electrode is (Figure 2.3)


Va =
2π a

And the absolute potential at a distance r is


Vr = (2.4)
2π r

Figure 2.4 gives the potential distribution around the electrode.

Rg = ρ/2πa

a r

FIGURE 2.2 Resistance to ground of a hemispherical electrode.

I
0
a

Vr = Iρ/2πr

Va = Iρ/2πa

FIGURE 2.3 Absolute potential at the electrode (Va) and at a distance r (Vr).
Resistance to Ground 9

Vr

V12

1 2 r

FIGURE 2.4 Potential distribution around a hemispherical electrode.

In order to determine the influence of the burial depth, we shall assume that the
conductor which carries the current to the electrode is insulated from ground. If a
spherical electrode of radius a is buried in a medium of infinite extent and resistivity
ρ Ω·m (Figure 2.5a), it is apparent that in this case

ρ ρI ρI
Rg = ; Vr = ; Vmax = (2.5)
4πa 4πr 4 ρa

All these values are one half the corresponding ones for a hemispherical electrode. If
the electrode is buried at a depth h meters below the surface of the ground as shown
in Figure 2.5b, we can determine its resistance and the potential and field distribution
at the earth’s surface as follows.
The direction of the current at the earth’s surface must be tangential to the surface
and this boundary condition is satisfied if we assume that the presence of an image
of the buried electrode at a vertical distance 2h from its center, i.e., at a distance h
above the earth’s surface as shown in Figure 2.5b and that the current entering the
image electrode is equal to that entering the real electrode. If we assume that 2h >>
a the potential at the surface of any one of the electrodes is

h r

x P

h r

(a) (b)

FIGURE 2.5 (a) Electrode buried in infinite medium. (b) Electrode buried at a depth h from
the surface.
10 An Introduction to Safety Grounding

Iρ Iρ Iρ  a 
Vo = + = 1+ (2.6)
4πa 4π(2h) 4πa  2h 

The resistance to ground of the buried sphere is therefore

ρ  a 
Rg = Vo /I = 1+
4πa  2h 


ρ 1  a 
=   1 +   (2.7)
2πa  2  2h 

and the potential of the point P on the ground surface is

Iρ  1 1  Iρ
Vx =  +  =
4π r r 2πr

Iρ (2.8)
=
2πa( x + h 2 )1/ 2
2

The maximum value of this potential (at x = 0) is

Iρ Iρ  a 
Vmax = =   (2.9)
2πh 2πa  h 

Figure 2.6a shows the potential distribution on the surface of the ground surround-
ing the buried electrode. It is evident that burying the electrode decreases the
potential difference which appears between any two points on the earth’s surface.

Vx
Ex
potential Vx of a
hemisphere at
ground level
potential Vx of
buried sphere

x
0.7h
h conductor isolated
h
from ground

(a) (b)

FIGURE 2.6  (a) Potential and (b) electric field distribution on surface of ground around
buried electrode.
Resistance to Ground 11

This is the potential difference which appears between the feet of a person walking


in the region around the electrode during the passage of a fault current to ground and
is referred to as the step voltage which will be discussed in detail in Chapter 4.
The potential difference between anybody connected to the ground electrode and
a point on the ground surface immediately above the electrode is

ρI  1 3 
Vo − Vmax =  −  (2.10)
4π  a 2h 

This potential difference is referred to as the touch voltage, and increases with
increasing burial depth.
The electric field at the point P is, from Eq.(2.8),

dV Iρ x
E x =− = (2.11)
dx 2π( x 2 + h 2 )3/ 2

It is zero at a point vertically above the electrode (x = 0) and has a maximum value
at a distance x = 0.7h:


Emax = (2.12)
2π(1.6h)2

Figure 2.6b shows the field distribution on the ground surface surrounding the buried
electrode.

2.2 MEASUREMENT OF ELECTRODE RESISTANCE TO GROUND


The most reliable and most frequently used method for measuring the resistance
to ground of an earth electrode is the so-called fall-of-potential method. In this
method (Figure 2.7), E represents the earth electrode and C and P are auxiliary
electrodes. If a current I flows between C and E and the potential difference
between E and P is V, then (subject to the conditions stated below) the ratio V/I
gives the required resistance to ground of the electrode E. The source of the current
I is the supply S which generates a constant voltage which should be alternating
to avoid any electrolytic action and its frequency should be higher than the power
frequency, between 70 and 80 Hz, to facilitate the elimination of power-frequency
stray currents.
If a series of measurements are made with different values of the distance H a
curve similar to that shown in Figure 2.7 is obtained. That part of the curve that is
quasi horizontal is the true resistance of the electrode E. As will be shown below
the extent of this horizontal section depends on the distance D between electrodes
E and C, i.e., on the degree of overlap of the resistance area (see Section 3.6) of
these electrodes.
Assume that the electrode E is replaced by an equivalent hemispherical electrode
of radius a (see Section 3.1). The current electrode C and the potential electrode P are
at distances D and H, respectively, from the center of the hemisphere. Suppose that
12 An Introduction to Safety Grounding

S
ac source A

I
E P C

D
H
resistance

probe position, H

FIGURE 2.7 Fall-of-potential method.

the current I enters at E and leaves at C. With reference to the previous section, it is
possible to express the potential at different points as follows:
Absolute potential at E due to current entering:

Iρ /2πa

Absolute potential at E due to current leaving at C:

−Iρ /2π(D − a) = − Iρ /2πD (D >> a)

The absolute potential at point E is therefore

Iρ  1 1 
VE =  − 
2π  a D 
Resistance to Ground 13

Absolute potential at P due to current entering at E:

Iρ /2πH

Absolute potential at P due to current leaving at C:

−Iρ /2π (D − H)

The absolute potential at point P is therefore

Iρ  1 1 
VP =  − 
2π  H D − H 

and the potential difference between points E and P is

VEP = VE – VP = V

Iρ  1 1 1 1 
=  − − + 
2π  a D H D − H 

If we let

D /a = c; H /a = p

the measured resistance between E and P is given by

ρ  1 1 1 
REP = V /I = 1− − +
2πa  c p c − p 

Since the actual resistance of a hemispherical electrode is ρ/2πa then the ratio
between the measured resistance and the actual resistance is

Measured resistance 1 1 1 
= 1−  + − (2.13)
Actual resistance  c p c − p 

The quantity between brackets represents the fractional error which results from the
choice of the distances D and H. The condition for zero error is

1 1 1
+ − =0
c p c− p

p2 + cp – c 2 = 0
14 An Introduction to Safety Grounding

the positive root of p is

p = 1 2 c( 5 − 1) = 0.618c

that is,

H = 0.618 D (2.14)

This result shows that whatever the distance D between the electrode E and C it
is possible to obtain the resistance of electrode E to ground when the distance H
between it and the potential electrode P is 61.8% of D.
Figure 2.8 shows a graphical representation of Eq. (2.13). The fractional error in
the resistance of electrode E is plotted as a function of the ratio p = H/a for different
values of the ratio c = D/a. It is apparent that the larger the value of c the smaller
the error in the actual value of the resistance of electrode E arising from any error
in ­setting the distance H = 0.618D. It has been agreed that the error in the actual
­resistance should be within ±2%.
Figure 2.9 shows the relationship between the radius a of the hemisphere equivalent
to the ground electrode and the distances D and H such that the error lies within ±2%.
When the earthing system consists of a number of electrodes or of a ground
­network the auxiliary electrodes must be placed outside the area of the earthing
­system. The potential electrode P should be located at a distance not less than 5 times

2.0

1.8

1.6
measured resistance/actual resistance

1.4

1.2

1.0
error limits ± %2
0.8

0.6

0.4

0.2

0
0 10 20 30 40 50
p

FIGURE 2.8  Graphical representation of Eq. (2.13).


Resistance to Ground 15

Distance Hor D (m)

Distance H or D (m)

Radius of equivalent hemisphere (m)

FIGURE 2.9  Separation of auxiliary electrodes to keep measuring error within ±2%.
(Curves adapted from Ref. [1].)

the longest radial distance of the area and the current electrode C at a distance not less
than 30m from P. In this case the distances D and H are measured from the e­ lectrical
center of gravity of the grounding system as shown in Figure 2.10. In this case

H ′ = 0.62 D′

( H + X ) = 0.62( D + X )

H = 0.62 D – 0.38 X  (2.15)

point of connection
with ground system

centre of gravity of ground system

E P C

X H
H'
D
D'

FIGURE 2.10  Measurement of the resistance to ground of an extended grounding system.


16 An Introduction to Safety Grounding

angle tower

C
P
θ
P
C ½θ

FIGURE 2.11 Positioning of auxiliary electrodes for measuring resistance to ground of an


angle tower.

The approximate value of X may be determined from the equation

X= A/π (2.16)

where A is the area which encloses the ground electrodes.


When measuring the resistance to ground of any overhead transmission line tower,
the auxiliary electrodes should be placed in a direction perpendicular to the direction
of the line in order to avoid any error resulting from induction effects. If the tower is
an angle tower (Figure 2.11) the auxiliary electrodes should lie along the line bisect-
ing the angle between the two line directions.

2.3 SOIL RESISTIVITY


Knowledge of the resistivity of the soil is of prime importance since the resistance to
ground of any buried electrode or system of electrodes is directly proportional to the
soil resistivity. The majority of soils and rock do not conduct electricity if they are
completely dry. However, if they contain moisture their resistivity decreases greatly
and they may be considered as conducting although their conductivity is very small
compared with that of metals. As an example the resistivity of copper is 0.017 µΩ·m
whereas that of ordinary soil is 100 Ω·m.1
The most important factors which determine the resistivity of a soil are as follows:

1. Type of soil
2. Moisture content
3. Types of salt dissolved and their concentration

5. Size of particles

1 The unit of resistivity in ohm-meter represents the resistance of a cube of side 1 m1 Ω·m= 100 Ω·cm
Resistance to Ground 17

The following table gives the approximate range for the resistivity of different
types of soils:

Sandy clay (mixture of clay, sand, and ash) 5–50 Ω·m


Loam 8–50
Mixture of sandy clay and stone 40–250
Sand and gravel 60–100
Sandstone and slate 10–500
Crystalline rock 200–10,000
Concrete (1 part cement + 3 parts sand) 50–300
Concrete (1 part cement + 5 parts gravel) 100–8,000

One of the most important of the factors which influence the soil resistivity is
the quantity of water retained in the soil, i.e., its moisture content, and the resistiv-
ity of this water itself and thus the type and concentration of salts dissolved in it.
The moisture content of any soil changes with changes in the weather, in the seasons,
in the nature of the subsoil, and in the depth of the water table. Except in deserts, it
is rare for a soil to be completely dry; however, it is also rare for the moisture content
to exceed 40%. In general the moisture content varies from about 10% in the dry
seasons to about 35% in the rainy seasons.
Measurements indicate that the value of the soil resistivity does not change much if
the moisture content exceeds 20%, but changes considerably as the moisture content
drops below that value (Figure 2.12). For example, we find that for a moisture content
of 10% the resistivity is 30 times that for a moisture content of 20%. Because of such
large variations, measurements of soil resistivity for grounding purposes should be
carried out in the dry season in order to represent the worst possible conditions as far
as resistance to ground of the earthing system is concerned.

4000
resistivity, ohm-meter

3000

2000

1000

0
0 10 20 30 40 50

moisture content %

FIGURE 2.12 Variation of the resistivity of clay soil with moisture content.
18 An Introduction to Safety Grounding

There is a considerable increase in soil resistivity as the temperature drops below


the freezing point. For example the resistivity of sandy clay with about 18% mois-
ture content rises from about 50 Ω·m at 20°C to about 360 Ω·m at – 15°C. This must
be taken into consideration when installing ground electrodes in cold regions; the
increase in resistivity can be minimized by burying electrodes below the frost line.
It should be pointed out here that in general the resistivity of a soil is associated
with its degree of corrosiveness. This is because resistivity is determined mainly
by moisture content as well as type and concentration of dissolved salts which such
moisture always contains. The following is a rough guide to the correlation between
resistivity and corrosiveness:

Resistivity (Ω·m) Degree of Corrosiveness


<5 Highly corrosive
5–10 Corrosive
10–20 Moderately corrosive
20–100 Mildly corrosive
>100 Slightly

When measurement of the soil resistivity at a certain site is difficult the following
values may be used as a guide:

Type of Soil Resistivity (Ω·m)


Moist organic soil 10
Moist soil 100
Dry soil 1,000
Rocky soil 10,000

2.4 
MEASUREMENT OF SOIL RESISTIVITY
There are essentially three principal methods of measuring soil resistivity:

• Simple box method


• The Wenner method
• Electromagnetic induction device

2.4.1 Simple Box Method


In this method a sample of soil from the site is well packed into a rectangular box
of insulating material with two identical plate electrodes at each end (Figure 2.13).
An AC source (not DC to avoid electrode polarization) is applied between the plates
and the current measured. The resistivity is simply obtained from Ohm’s law.
This simple method is suitable for testing in situ if it is known that the soil ­consistency
at the site is uniform over a large area and sufficient depth. Otherwise it is a cheap
Resistance to Ground 19

= / = / L Area of end
plates (S)
ρ = VS/IL

FIGURE 2.13 Simple box for measurement of soil resistivity.

and simple laboratory method to determine the resistivity of different types of soil
as well as the effect which moisture, temperature, and additives (such as salts) have
on soil resistivity.

2.4.2 Wenner method


This is the most common method of measuring the soil resistivity (Figure 2.14).
In this method four electrodes are driven into the ground along a straight line at equal
intervals. A known current I passes between electrodes 1 and 4 and the potential
difference V is measured between electrodes 2 and 3. Assuming that the probe length
is small compared to electrode separation (which is the usual case) we have that
Absolute potential at 2 due to current entering at 1:

Iρ /2πL

Absolute potential at 2 due to current leaving at 4:

−Iρ /4πL

The total absolute potential at 2 is therefore

V2 = Iρ /4πL

V
V
I I

1 2 3 4

L L L

FIGURE 2.14 Wenner method of measuring soil resistivity.


20 An Introduction to Safety Grounding

Similarly the total absolute potential at 3 is

V3 = −Iρ /4πL

Hence the potential difference between 2 and 3 is

V23 = V2 – V3 = V = Iρ / 2πL

and the soil resistivity is given by

ρ = (V /I)2πL = 2πLR (2.17)

where ρ is the actual soil resistivity if the soil is homogeneous. However, since
the resistivity of a soil usually varies with depth (top-soil layer, subsoil layer), the
measured resistivity ρ represents the apparent resistivity of the soil at the measur-
ing site. By increasing the probe spacing L, the current penetrates a deeper and
longer distance and gives a better estimate of the resistivity if the latter varies
appreciably with depth.
If the length of the probe l is not small compared with the probe separation L, then
the apparent resistivity is given by

4πLR
ρ= (2.18)
2L L
1+ −
L2 + 4l 2 L2 + l 2

If L ≫l then ρ = 2π LR.
When measuring the soil resistivity at a site it is recommended to carry out several
measurements with different values of the distance L between the electrodes to obtain
an average value for the resistivity. Any change in the resistivity with changes in L
is an indication that the soil is not homogeneous and in particular that the resistivity
changes with depth. This is because the greater the distance between the electrodes
the greater the depth to which the current penetrates. In such cases it is a common
practice to assume that the apparent resistivity measured with the electrodes at a
distance L apart is the average resistivity of the soil up to a depth L. Although such
an assumption is not accurate it is accepted from the practical point of view.

2.4.3 electromagnetic induction method


This method was developed primarily for the measurement of soil conductivity for
geophysical prospecting. It consists essentially of two coils: a transmitter coil T fed
by an audio frequency generator (10–20 kHz) and a receiver coil R at a distance
s away from T. To measure the resistivity for grounding purposes both coils are
placed with their plane flat over the ground surface (vertical polarization position)
as shown in Figure 2.15. The varying magnetic field produced by the transmission
coil T induces eddy currents in the soil beneath. These currents flow in closed loops
Resistance to Ground 21

s
Transmitter coil T Receiver coil R

Induced eddy
currents

FIGURE 2.15 Principle of electromagnetic induction for measuring soil conductivity


(resistivity).

perpendicular to the magnetic field. The magnitude of the currents is proportional


to the magnitude of the field, to the rate of change of this field and to the area of the
loop, and inversely proportional to the resistivity of the medium in which they flow.
The eddy current loops generate their own magnetic field and the receiver coil senses
both this field and the field of the primary coil at the receiver location.
Provided that certain design criteria are met2 the ratio of the secondary to the
primary fields is linearly proportional to the apparent soil conductivity and can be
expressed as

1 4  Hs 
=σ = (2.19)
ρ ωµ0 s 2  H p 

where
ρ = Ω

s = intercoil spacing (m)


Hs = secondary field at receiver loop
Hp = primary field at receiver loop

This linear relationship is valid in the range 1–1,000 Ω·m.


It should be mentioned that measurements can be made with both coils placed
either flat on the ground (vertical polarization) or with their planes perpendicular to
the ground (horizontal polarization). In the first position the depth of penetration is
about twice greater than in the second position and should be used when measuring ρ
for grounding purposes. However, measurements made with horizontal polarization
provide information on the resistivity of the soil layer near the surface.

2 J.D.McNeill, Electromagnetic Terrain Conductivity Measurements at Low Induction Numbers,


Technical Note, Geonics Ltd., 1980.
22 An Introduction to Safety Grounding

The great advantage of the electromagnetic induction method (EM) is that it is fast
so that measurements over large tracts can be accomplished quickly. Another advan-
tage is that since there are no driven rods it avoids current injection problems (gravel,
bedrock, snow, ice, etc.) and is ideal for measurements in terrain where the driving of
rods can be very difficult (e.g., rocky ground). Results obtained by this method com-
pare well with those obtained by the Wenner method. However, EM equipment is about
10–20 times more expensive than that required for a Wenner test. Commercial EM
measuring instruments are available for exploration depths from 6 to 30m.
Whatever the method of measurement used, a record should always be kept of
the date the measurements were made, the condition of the soil (wet or dry), the
temperature, and any information regarding the presence (or suspected presence) of
bare conductors buried at the site. The presence of such conductors can greatly affect
the path of the current flow in the ground and hence the value of the soil resistivity.

2.5 COMPUTER PROGRAMS


The existence of more than one formula for the resistance to ground of most electrode
geometries indicates that the formulas are not exact but have nevertheless proved to
be very good approximations. In fact they are obtained by mathematical approxima-
tions. The choice of the approximation method depends essentially on the geometric
configuration of the electrode or electrode system, on the boundary conditions, and
on the degree of accuracy required. Numerical methods are best since they can be
readily computerized. The two most common of these methods best suited for solv-
ing the partial differential equation which governs the system’s behavior (e.g., the
Laplace equation∇ 2V=0) are the finite difference method and the finite element (FE)
method. Of these the FE method is better suited to systems with irregular geometry,
unusual boundary conditions, or nonhomogeneous media.
Essentially the FE method consists of subdividing the region of interest into a num-
ber of finite simply shaped elements which could be one-, two-, or three-dimensional.
The finer the subdivisions the more closely will the resulting FE structure represent the
original configuration. Nodes are the points where elements “connect” and the collection
of nodes constitutes the FE mesh. An approximate solution is then developed for each
element and the total solution is obtained by assembling the individual solutions such that
all boundary conditions are satisfied at all interelement boundaries. Thus the partial dif-
ferential equation is solved in a piecewise manner. It should be pointed out that the total
solution requires the solution of a very large number of simultaneous linear equations.
This is such an extremely laborious and time-consuming task that FE methods were only
used in simple cases. However, with the advent of powerful and fast computers, FE has
become the method of choice for the determination (among its many other applications)
of the resistance to ground of a large variety of electrode geometries and configurations
(e.g., grounding grids) in both uniform and non-uniform soils. There are a number of
commercially available programs such as ETAP, CYMGRD, or CONSOL Multiphysics
which use the FE method as an alternative with automatic mesh generation.
It should be pointed out that whatever the method used to predetermine the value
of the resistance to ground of a chosen configuration, the actual value can only be
obtained by direct measurements after the system has been installed.
3 Earthing Electrodes

3.1 SINGLE DRIVEN ROD


Although the hemispherical electrode is the simplest geometric shape for calculating
its resistance to ground, from the practical point of view, the most common type of
earthing electrode (especially in distribution networks) is the solid cylindrical rod
driven vertically into the ground (Figure 3.1).
A fairly accurate expression for the resistance of a rod electrode is given by
Laurent1

ρ 3L
R= ln (3.1)
2πL d
where ρ is the soil resistivity, d is the rod diameter, and L is the length of the rod.
Other formulas frequently used are that given by Tagg,2

ρ 4L
R= ln (3.2)
2πL d
and that given by Dwight,3

ρ  8L 
R=  ln − 1 (3.3)
2πL  d 

FIGURE 3.1 Driven rod electrode.

1 P. Laurent, General fundamentals of electrical grounding techniques, in IEEE Std 80-1986(R1991):


Guide for safety in AC substation grounding.
2 G.E. Tagg, Earth Resistances, Newnes, London, 1964.
3 H.B. Dwight, Calculations of resistances to ground, Elec. Engr., Vol 55, p 1319, 1936.

23
24 An Introduction to Safety Grounding

As a typical example if we assume that d = 5 cm, L = 6 m, and ρ = 100 Ω·m, we find


that:
From Eq.(3.1),R = 15.60 Ω
From Eq.(3.2),R = 16.17 Ω
From Eq.(3.3),R = 15.55 Ω
The differences between these values are of no practical significance especially
in view of the inherent inaccuracy in the value of the soil resistivity ρ. In the present
text we shall use Eq. (3.1).
By equating Eq.(3.1) with Eq. (2.2) we can obtain the radius a of the equivalent
hemispherical electrode, i.e., the hemisphere which has the same resistance to ground
as the rod,

L
a= (3.4)
ln(3L /d )
Because the rod diameter d appears in the logarithmic term in Eq. (3.1), the magnitude
of the resistance to ground of a rod electrode does not change significantly with its
diameter and we can therefore consider that its resistance is directly proportional to
the earth resistivity and inversely proportional to the length of the rod. On the other
hand, the resistance to ground of a hemispherical electrode is directly proportional
to the earth resistivity and inversely proportional to its radius. If we assume the
resistivity increases 100 times then, for a given resistance to ground, the length of a
rod would have to be increased 100 times. The radius of the hemisphere would also
have to be increased 100 times and hence the excavated volume would increase a
million times. This example shows the principal advantage of using driven rods as
ground electrodes.
Table 3.1 gives the variation of resistance with rod length for two rod diameters.
It is apparent that the rod diameter has no significant effect on the rod resistance to
ground. In practice the diameter of the rod is chosen such that the rod can withstand
being driven into the ground without bending or any other mechanical damage.
Figure 3.2 shows the relationship between rod length and its resistance to ground for
d = 2.5 cm and ρ = 100 Ω·m.

TABLE 3.1
Variation of Resistance to Ground with Rod Length
Resistance Resistance Equivalentradius
Length R(Ω) R(Ω) a(m)
L(m) d = 5cm d = 2.5cm d =2.5cm
1 0.61 ρ 0.76 ρ 0.21
2 0.38 ρ 0.44 ρ 0.36
4 0.22 ρ 0.25 ρ 0.64
8 0.123ρ 0.137ρ 1.16
16 0.07 ρ 0.075ρ 2.12
32 0.04 ρ 0.04 ρ 3.98
Earthing Electrodes 25

80

70

resistance R(ohm)
60

50

40

0 3 6 9 12
rod length L(m)

FIGURE 3.2 Relationship between R and L for a driven rod.

3.2 GROUNDING USING MULTIPLE RODS


In many cases it is found that extending the length of a single rod is not enough to
attain the required value of resistance to ground especially when the soil resistivity is
high. In such cases a number of rods connected in parallel must be used.
When a number of electrodes in parallel are used, the resistance to ground is not
reduced by the same number unless the distance between the electrodes is infinite.
Since this is not practical, it is necessary to know what effect the distance separat-
ing the electrodes has on the reduction of the resistance of a single electrode. In the
following we will calculate this reduction for two electrodes and for three electrodes
in a straight line.

(a) Two rods in parallel

Suppose that we have two identical rods at a distance d apart and that the radius of
the equivalent hemisphere is a (Figure 3.3a). Because of symmetry each electrode
will carry a current of ½ I to ground. With reference to Section 2.1 we may write the
absolute potential of any one of the electrodes as

I I
½I ½I I1 I1
I2
a a

d d d

(a) (b)

FIGURE 3.3 (a) Two rods in parallel; (b) three rods in parallel.
26 An Introduction to Safety Grounding

ρ 1 1  1
V= + 2 I
2π  a d 
and the total resistance of the two electrodes in parallel is

ρ
R = V /I = 1
2 [1 + (a /d)]
2πa
If we let α = a/d we may write,

Resistance of two rods in parallel 1


= 2 (1 + α ) (3.5)
Resistance of one rod
The above ratio will be referred to as the resistance ratio. It is evident that this ratio
does not equal 0.5 except if the separation d is infinite.

(b) Three rods in parallel

We shall assume that the three rods are along a straight line as shown in Figure 3.3b.
Because of symmetry the currents flowing in the outer electrodes are equal. If I1 is
the value of each of these currents and I2 is the current in the central electrode then,

I = 2I1 + I 2 (3.6)

and the absolute potential of any of the two outer electrodes (assuming that d >> a) is

I1 ρ I 2 ρ I1 ρ
V= + +
2πa 2πd 2π(2d)
ρ
= I1 + α I 2 + 12 α I1 
2πa 
ρ
= I1 (1 + 12 α ) + α I 2 
2πa 
and the potential of the central electrode is

ρ
V= [ I 2 + 2α I1 ]
2πa
Since these two potentials are equal,

I1 (1 + 12 α − 2α ) = I 2 (1 − α )

I1 (1 − 32 α )
I2 = = kI1
(1 − α )
(3.7)
(2 − 3α )
k=
2(1 − α )
Earthing Electrodes 27

From Eqs. (3.6) and (3.7) we find that

I
I1 =
2+k
and the potential V can be written as

ρ  kI 2α I 
V= +
2πa  2 + k 2 + k 


ρ I  k + 2α 
=
2πa  2 + k 
Substituting for the value of k we find that the resistance ratio is

(2 − 3α )
+ 2α
2(1 − α ) 2 − 3α + 4α − 4α 2 2 + α − 4α 2
= =
(2 − 3α ) 4 − 4α + 2 − 3α 6 − 7α
2+
2(1 − α )

Table 3.2 shows the resistance ratio for a number of electrode configurations and
Figure 3.4 shows a set of curves from which the resistance ratio can be obtained for
these configurations. It has been found that the calculated values of this ratio are
always between 5% and 20% higher than the measured value and is therefore always
on the safe side.
When the number of rods required to obtain a certain value for the resistance
to ground increases, the most common configurations used are the following
(Figure 3.5):

In many cases the area available for grounding purposes is limited. In such cases
it is important to know the number and configuration of rods which will give the
optimum utilization of the available area in order to obtain an effective grounding

TABLE 3.2
Resistance Ratio of Different Rod Configurations
Configuration Resistance Ratio
Three rods in a straight line (2 + α − 4α2/(6−7α)
Three rods in equilateral triangle (1+2α)/3
Four rods in a straight line (12+16α−21α2)/(48−40α)
28 An Introduction to Safety Grounding

resistance ratio

distance between rods,cm

FIGURE 3.4 Relationship between resistance ratio and distance between rods for a number
of configurations. A, two rods; B, three rod in a straight line; and C, four rods in a straight
line. (Adapted from Ref. [1].)

(a) (b)
hollow solid

(c)
circle

FIGURE 3.5 Rod configurations.

system. Figure 3.6 shows the relationship between the resistance to ground and the
number of rods in a hollow square arrangement required to attain this value within
different areas. Figure 3.6 indicates that there is a lower limit to the value of the resis-
tance to ground which may be attained and also an economic limit to the number of
rod which must be used to obtain this resistance. For example, for an area of 36 m2,
we find that the lower limit for the resistance is 6 Ω and that the economic number
of rods is 16. If a resistance to ground of less than 6 Ω is required then a larger area
Earthing Electrodes 29

resistance (ohms)

number of rods

FIGURE 3.6 Resistance to ground of rods arranged in a hollow square of finite area.
(Adapted from Ref. [1].) Rod length, 2.4 m; rod diameter, 2.5 cm; ρ = 100 Ω·m.
resistance, ohms

hollow square
solid square

area, sq.m

FIGURE 3.7 Minimum resistance obtainable with hollow and solid square rod configura-
tions in a limited area [1]. Rod length, 2.4 m; rod diameter, 2.5 cm; ρ = 100 Ω·m.

must be used. If the earth resistivity was 2000 Ω·m instead of 100 Ω·m the lower
resistance limit would be 120 Ω (6 × 2000/100) instead of 6 Ω.
Figure 3.7 shows the relationship between area and lower resistance limit for
both hollow square and solid square rod configurations. It is evident that if the
area is limited then the addition of rods inside a hollow square produces only a
very small decrease in resistance value which makes the addition of such rods
uneconomical.
30 An Introduction to Safety Grounding

3.3 GROUNDING USING HORIZONTALLY BURIED WIRES


If the nature of the ground is rocky or if there is a layer of rock near the surface so
that it is difficult to use driven rod electrodes, then a horizontal wire buried at a depth
of from 0.5 to 1 m is used as a ground electrode. This wire can have different shapes:
straight line, right angle, or multiarm star (Figure 3.8).
Figure 3.9 shows the variation of the resistance to ground with the length of a
straight horizontal buried wire. It is apparent that if the length of the wire exceeds
90 m the rate of decrease of resistance with wire length becomes very small.
In general the resistance to ground of a buried wire electrode may be found from
the following equation:

ρ
R=
2πL
[ ln(4l /d ) + f (h /l )] (3.8)

where
L = total length of the wire
l = length of one arm
d = wire diameter
h = depth of burial

l
l l L

FIGURE 3.8 Common shapes of buried wire.


resistance, ohms

wire length,m

FIGURE 3.9 Variation of resistance to ground with wire length for a horizontal wire buried
at a depth of 60 cm (wire diameter, 2.5 cm; ρ = 100 Ω·m).
Earthing Electrodes 31

The function f (h/l) is a depth factor whose value depends on the wire configuration.
Its value may be obtained from the curves given in Figure 3.10 for six different wire
shapes [1]. As an example if L = 150 m, d = 2.5 cm, h = 90 cm, ρ = 100 Ω·m, the
resistance values to ground of the different wire shapes are as follows:

Straight wire ………………….1.41Ω


Right angle ………………….. 1.44 Ω
3-arm star …………………….1.50 Ω
4-arm star …………………….1.62 Ω
6-arm star …………………….1.94 Ω
8-arm star …………………….2.33 Ω

The above figures show that for a fixed length of wire the straight wire gives the least
resistance. If we now assume that we have a limited area whose diameter is say 60 m,
then for the same above values of d, h, and ρ, the resistance, the arm length and the
total wire length for the different wire configurations which can be accommodated
within the given area are as shown in Table 3.3.
f (h/l)

1. Straight wire
2. Right angle
3. 3-arm star
4. 4-arm star
5. 6-arm star
6. 8-arm star

FIGURE 3.10 Depth factor f (h/l) for different configurations of buried wire [1].
32 An Introduction to Safety Grounding

TABLE 3.3
Resistance to Ground of Different Wire Shapes
Configuration R (Ω) l (m) L (m)
Straight wire 3.0 60 60
Right angle 2,26 42.4 84.8
3-arm star 2.31 30 90
4-arm star 1.99 30 120
6-arm star 1.63 30 180
8-arm star 1.49 30 240

Area diameter 60 m, d = 2.5 cm, h = 90 cm, ρ = 100 Ω·m.

From this Table 3.3 we see that the resistance of an 8-arm star is half that of a
straight wire. However, because a very large increase in wire length is required to
achieve only a small decrease in resistance, the common practice is to limit the num-
ber of star arms to a maximum of four.
It should be mentioned here that the straight wire could very well be a buried
water metal pipe; its resistance to ground can be estimated as above and then prefer-
ably checked by measurement.
An alternative to straight and star wires is to use a wire loop around the periph-
ery of a structure, buried at a minimum depth of 0.5m and about 1m away from the
external walls of the structure. This is known as a ring earth electrode and is used as
an alternative method for earthing lightning down conductors.

3.4 GROUNDING USING BURIED PLATES


Buried plates were extensively used in the past as grounding electrodes, but because
of the poor efficiency of metal usage (i.e., the amount of metal required to attain
a given resistance) compared with that of driven rods or buried wires, their use is
uneconomical.
The resistance to ground of a circular plate of radius a with its center at a distance
h below the ground and buried either horizontally or vertically (Figure 3.11) can be
obtained from the following equation:

ρ
R= [1 + a /(2.5h + a)] (3.9)

and for large depths,

ρ
R= (3.10)

It should be mentioned that the thickness of the plate has no significant effect on its
resistance to ground.
Earthing Electrodes 33

a a

FIGURE 3.11 Buried circular plate.

3.5 WATER PIPES OR STEEL REINFORCEMENT


AS GROUNDING ELECTRODES
It is sometimes possible to use the metal pipes of a water network as earthing
electrodes. However, before such a decision is taken, the network’s resistance to
ground must be measured and the electrical continuity of all the joints must be
ensured (by means of jumpers). The electrical continuity must be checked after any
repair or maintenance work is carried out on the system. If the replacement of metal
pipe section by plastic pipes is envisaged or if the electrical continuity cannot be
guaranteed, then this method should not be used.
In buildings where steel reinforcement is used in the concrete foundations such
as column bases, the reinforcement bars can be used as grounding electrodes. The
resistivity of the concrete below the surface of the earth is about 30 Ω·m which
is lower than the average resistivity of the soil itself (100 Ω·m). In such cases it is
sufficient to make the connections between any main reinforcing rod and the main
earth connection at the distribution board.

3.6 THE RESISTANCE AREA


Figure 2.2 shows that as the distance from the grounding electrode increases, the
ratio between the value of the resistance up to this distance and the total (absolute)
resistance of the electrode increases. It is clear that the rate of this increase is rapid
at first but then becomes asymptotic. Theoretically the total resistance to ground
can only be attained at infinite distances. For practical purposes, we can say that
surrounding any electrode there is a finite region of ground that contains the “major
part” of the resistance. There is general agreement that this major part should repre-
sent 98% of the total resistance and it is known as the resistance area.
As shown in Chapter 2 the resistance between a hemispherical electrode of radius
a and a point at a distance r from its center is

ρ  1 1
R=  − 
2π  a r 

The ratio x of this resistance to the total resistance ρ/2πa, expressed as a


percentage, is
34 An Introduction to Safety Grounding

x = 100 (1 – a /r )

so that

100a
r= (3.11)
100 − x
If we assume that the grounding electrode is a rod of length 2.5 m (which is the
minimum length permitted) and diameter 2.5 cm, then the radius of the equivalent
sphere from Eq.(3.4) is 43.2 cm and the radius of the resistance area is

r = 43.2 / (100 – 98 ) = 21.6 m

If it is required to obtain an effective electrical separation between two different


grounding electrodes then the distance between them should not be less than 20 m.
If the distance is less then there is an overlap of their respective resistance areas.
Electrically this means that there is a mutual resistance between the two grounding
points (Figure 3.12a). The effect of this resistance may be clarified from Figure 3.12b;
when a short circuit occurs between one phase and the body of the transformer, the
potential of the point E rises to a value which depends on the magnitude of the resis-
tance to ground at E and that of the short circuit current. The potential of the point S,
and hence of the body grounded through S, rises to a value determined by the value
of the mutual resistance Rm between points S and E. The closer the points the higher
the mutual resistance.

overlap of
resistance areas

E S

(a)

R1 R2

Rm
(b)

FIGURE 3.12 Overlap of resistance areas.


4 Step, Touch, and
Transfer Voltages

4.1 STEP VOLTAGE


If a man is standing with his feet apart or walking in an area in the vicinity of a
ground electrode (in particular a substation ground or a short-circuit testing facility
ground) and a short-circuit current I flows to ground then a voltage, known as the
step voltage, will appear across his feet. Suppose that the man is at a distance x from
the electrode and that the distance between his two feet is s. If the ground electrode
is represented by an equivalent hemisphere of radius a, then the absolute potential at
a distance x from its center is (from Eq.2.4)

Vx =
2πx
and the potential difference appearing between the man’s feet is therefore

Iρ  1 1  Iρ  s 
Vstep =  − =
2π  x x + s  2π  x ( x + s) 

It is evident that the maximum value of this potential occurs at x = a. If we assume


that the length of the step s is one meter, the maximum step voltage is

Iρ  1 
Vstep (max.actual) =
2π  a(a + 1) 

If we assume that the resistance of the body is Rb and the resistance between each
foot and ground is Rf (Figure 4.1), then the current passing through the body is

I b = Vs / ( Rb + 2 R f )

Since the current which passes through the human body must not exceed 0.116/√t (see
Section 1.3), then the safe step voltage, i.e., the maximum step voltage permitted is
given by

Vstep ( max.permitted ) = I b ( Rb + 2 R f )
(4.1)
= (1,000 + 2 R f )0.116/ t

and it must therefore always be ensured that

Vstep ( max.permitted ) ≤ Vstep ( max.actual )

35
36 An Introduction to Safety Grounding

Rb

Rf Ib Rf

R2
Vs
I
Vs

Rb

Rf s Rf
hemisphere
of radius a

R1 R2

FIGURE 4.1 Step voltage.

4.2 TOUCH VOLTAGE


If a person touches a metal object directly connected to ground at the same instant as
a short circuit current is flowing to ground (Figure 4.2), the magnitude of the current
which flows between his hand and feet is determined by the so-called touch voltage.
Since I >> Ib, the touch voltage may be written as

Vtouch = IR1 = I b ( Rb + 1 2 R f )

Vtouch
I

I Ib Rb
Vtouch
R1
½R f

R2

Rb

½R f
hemisphere
of radius a
R1 R2

FIGURE 4.2 Touch voltage.


Step, Touch, and Transfer Voltages 37

If we assume that the person is at a distance of 1 m from the metal object which he
touches then from Eq. (2.1),

ρ 1 
R1 =  − 1
2π  a 
setting,

I b = 0.116 / t and Rb = 1,000 Ω

we find that the safe touch voltage, which is also the maximum permissible touch
voltage, is

Vtouch ( max.permitted ) = (1, 000 + 1 2 R f ) 0.116/ t (4.2)

It is apparent from Eqs. (4.1) and (4.2) that the magnitude of the resistance Rf between
foot and ground has a very great influence on the maximum permissible value of
both the step and touch voltages. For practical purposes it is possible to regard the
foot as a circular electrode whose radius is about 8 cm; the resistance to ground of
such an electrode is obtained from Eq.(3.9) for h = 0, a = 0.08,

ρ ρ
Rf = = ≅ 3ρ
4 a 4 × 0.08

where ρ is the resistivity of the soil. Substituting this value of Rf in Eqs. (4.1) and (4.2)
we find that maximum permissible step voltage is
(116 + 0.7ρ )
Vstep (max.permitted) = (4.3)
t
and that maximum permissible touch voltage is
(116 + 0.17ρ )
Vtouch (max.permitted) = (4.4)
t
Equations (4.3) and (4.4) are valid for uniform soil resistivity. In many cases, how-
ever, the resistivity is not uniform either because of the nature of the soil or because
the ground surface has been covered with a layer of high resistivity (ρs) such as
crushed rock. In this case these equations are modified as follows:

(116 + 0.7Cs ρs )
Vstep (max) = (4.5)
t

(116 + 0.17Cs ρs )
Vtouch (max) = (4.6)
t
where Cs = correction factor for derating the surface resistivity ρs.
The risk of electrocution is higher for touch voltages than for step voltages as in
the former the current traverses the chest region and its path is close to the heart.
38 An Introduction to Safety Grounding

It is evident from the last equations that the higher the surface resistivity the
higher the permissible step and touch voltages. This is very desirable at large substa-
tions where the ground fault current is generally high as well as around lightning
ground electrodes. In practice the surface resistivity is increased by covering the
surface with a layer of crushed rock such as granite of thickness 10–12 cm. Granite
has a very high resistivity when dry and also when wet: dry 1.3 × 106Ω·m, wet 4.5 ×
103Ω·m. If granite is not available then asphalt may be used as it too has a high wet
resistivity > 104Ω·m [4].
The sudden change in resistivity at the boundary between the two layers is
described by the reflection factor K defined as

ρ − ρs
K= (4.7)
ρ + ρs
Figure 4.3 shows the relationship between the correction factor Cs and the thickness
of the surface layer hs for different values of K (IEEE Std.80-2000).
Equations (4.5) and (4.6) must be used whenever a surface layer (crushed rock or
any other material) has been added or when the soil can be represented by two layers
(e.g., top soil + subsoil). It is also possible to obtain an approximate value for Cs from
the following formula:

 ρ
0.09  1 − 
 ρ s 
Cs = 1 − (4.8)
2hs + 0.09

1
0.9

0.8

0.7
Cs
0.6

0.5

0.4

0.3

0.2

0.1

0
0 0.04 0.08 0.12 0.16 0.2 0.24 0.28 0.3
Thickness of surface material, hs meters

FIGURE 4.3 Relation between correction factor Cs and thickness of surface layer hs.
(Reproduced with permission from IEEE Std. 80-2000.)
Step, Touch, and Transfer Voltages 39

4.3 TRANSFER TOUCH VOLTAGE


The mutual resistance between two separately earthed systems is, as mentioned in
Section 3.6, a source of transfer voltage. In general the transfer touch voltage can be
defined as the potential which appears between the grounded metallic parts of two
systems which are separately grounded when a fault to ground occurs on one of these
systems.1 This potential can be especially dangerous when one location is a power
substation, since under fault conditions the ground potential rise may be of the order
of several thousand volts. Two such cases are shown in Figure 4.4.

Vtransfer

separate ground
short circuit
current (Isc) Rb

Rf/2
R1 R2

(a)

touch
ground conductor, touch voltage
voltage safe
pipe or rail leaving dangerous
the station

short circuit
current (Isc)

ground mat at station

(b)

FIGURE 4.4 Transfer voltage.

1 Separate grounding at the power source and at the utilization (consumer) end is also the source of
common-mode noise (voltage and current). Treatment of this topic is beyond the scope of this book and
the interested reader may consult any specialized text on the subject, e.g. H.W.Denny, Grounding for
the Control of EMI, Don White Consultants, Inc.1983, USA.
40 An Introduction to Safety Grounding

In Figure 4.4a the absolute potential at the point where the man is standing during
the flow of a current I to the earth of a grounded system is Vx = I ρ/2 π x = I R2. If the
man comes in contact with a metallic body which is earthed through a second ground
system which is physically separated from the first, then this is the value of the touch
voltage he will experience.
In Figure 4.4b the man standing on the properly designed (see Chapter 6) substa-
tion ground mat is in no danger under ground fault conditions. However, if a man
standing outside the station were to touch a ground conductor or pipe or rail leaving
the station he may be in considerable danger.

4.4 GROUNDING OF POWER TOWERS


All power towers of overhead transmission lines must be individually grounded. The
resistance to ground, also known as the tower footing resistance, should not exceed
10 Ω for towers at a distance less than 2 km from the line ends and 20 Ω for all
other towers. The concrete foundations to which the tower feet are fixed can serve
as ground electrodes although it is preferable, in the majority of cases, to use one of
the grounding arrangements shown in Figure 4.5. In some cases it may be necessary
to use a large amount of buried wires, especially if the ground is rocky or the soil
resistivity very high because of extremely dry conditions.

4.4.1 effect of overhead ground Wire on toWer footing reSiStance


In order to protect overhead transmission lines against direct lightning strokes the
majority of lines have one or more ground wires mounted on top of the towers and
extending over the entire length of the line (Figure 4.6a). Any discharge to ground due
to lightning or flashover at an insulator will cause the flow of current to ground. This
current will be distributed between the ground wire and all the towers. Depending on
the tower footing resistance, ground wires and towers will be raised to a high poten-
tial above ground; thus a lightning current of 10 kA and a tower footing resistance of
10 Ω will raise the voltage to 100 kV. This voltage, in addition to the service voltage,
may cause a flashover between a conductor and tower causing a line-to-ground fault.
Figure 4.6b shows the equivalent “ground” network in which R represents the stand-
alone resistance to ground of each tower (this resistance may vary slightly between
towers but such variations are small and may be neglected) and r is the resistance
between two successive towers. These resistances form a ladder network.2
If we assume that the line is infinitely long in both directions, then the input resis-
tance of such a network between the points a and b (or c and d) can be shown to be

r
Rab = Rcd ' = R tanh (4.9)
R
and since always R >> r we have that

2 A.A. Zaky and T.T. El-Sonni, Simplified method for determining the resistance to ground of power
transmission towers, YJES, Vol.6, pp 1-5, 2013.
Step, Touch, and Transfer Voltages 41

ground wire for


protection against
lightning

additional buried wire


having the length necessary
≥1m
to obtain a resistance ≤10Ω

ground electrode
tower foot
foundation

ground ≥4m
conductor

buried loop
interconnecting
ground electrodes
(50mm 2 copper)

(a) (b)

FIGURE 4.5 Grounding of a power transmission line tower.

r r
tanh =
R R

Rab = Rr
42 An Introduction to Safety Grounding

(a)

If

r r c a r r

R R R R R
d b

(b)

If

If ʹ
(c) √(Rr) R √(Rr)

FIGURE 4.6 Ladder network formed by towers and ground wires.

From the equivalent circuit shown in Figure 4.6c we see that the total resistance to
ground consists of the resistance R of a single tower in parallel with the two resis-
tances Rab and Rcd; thus

R 12 Rr
Rg = ≅ 1
2 Rr (4.10)
R + 12 Rr

The current through the tower at which the fault occurs is

I f 12 Rr
I ′f = ≅ If 1
2 Rr (4.11)
R + 12 Rr

and hence

Rg I ′f r
= = 1
2 (4.12)
R If R

This result indicates that the presence of the ground wire reduces the current through
the tower and its resistance to ground by the same ratio. As a numerical example
assume that a long power line has a stranded steel ground wire of 10 mm diameter
whose resistance is 6.7 Ω/km, that the distance between successive towers is 180 m,
and that the footing resistance of each tower is 10 Ω. From Eq.(4.10) we find that
Step, Touch, and Transfer Voltages 43

Rg = 1
2 10 × 1.2 = 1.73 Ω

This shows that the presence of the ground wire leads to a large decrease in the
resistance to ground at each tower; this means that the step and touch potentials are
greatly reduced as is the danger of flashover. If the towers are close to the ends of the
line (less than 2 km) the resistance to ground is

Rg = Rr

which is double the resistance of the more distant towers.


5 Grounding Systems

5.1 MAGNITUDE OF RESISTANCE TO GROUND


The ideal value of the resistance to ground is zero. Although in practice this can
never be attained, it is possible to obtain values of less than 1 Ω. However, in many
cases such low values are not necessary. In general the value of the desirable resis-
tance is inversely proportional to the magnitude of the short-circuit current flowing
to earth, i.e., the higher the current the smaller the resistance. Since the reactance of
cables is lower than that of overhead lines, short-circuit currents are higher in cables
than in overhead lines so that the resistance to ground of installations fed by cables
must be lower than that of installations fed by overhead lines.
For generating stations and large substations the resistance to ground should not
exceed 1 Ω. For small substations and industrial installations the resistance should
be usually less than 5 Ω.
It should be mentioned that the standard specifications of many countries specify
the value of the maximum permissible resistance to ground. For example in the
USA the National Electric Code (NFPA 70: National Electrical Code) recommends
that the resistance should not exceed 25 Ω whereas the German Standard (VDE
0100) specifies 5Ω as the maximum value.

5.2 ELEMENTS OF THE GROUNDING SYSTEM


In general any grounding system consists of the following elements (Figures 5.1
and 5.2):

1. An area of land of suitable earth resistivity and in which earth electrodes


can be readily driven or buried.
2. The earthing electrodes.
3. Joints.
4. The grounding conductors which are the conductors used for connecting
the electrodes together and connecting the electrodes either to the ground
connection provisions or directly to the equipment and structures which are
to be grounded.
5. Ground connection provisions. In the case of small installations it is pos-
sible to use grounding conductors to connect the ground electrode directly
to the equipment to be grounded. However in the case of large installations
it is preferable to have special provisions consisting of ground bus-bars and
suitable terminals for grounding conductors and equipment grounding con-
nections. These provisions should be mounted on walls or foundations at loca-
tions which are easily accessible; this is necessary for inspection purposes.

45
46 An Introduction to Safety Grounding

ground connection provisions


(ground bus-bar)

grounding
equipment grounding conductor
conductor

ground electrode

FIGURE 5.1 Elements of a grounding system.

L
N
signal reference system
G (depends on type of equipment)

equipment
grounding
conductor lightning down
distribution
conductor
board

L1
L2
ground
L3 provisions
N

entrance
panel

grounding
conductors
water pipes

electrodes

FIGURE 5.2 Elements of a supply system (residential or commercial).


Grounding Systems 47

6. Equipment grounding conductors (EGC). These are the connections


between the noncurrent carrying parts of the equipment to be grounded and
the ground connection provisions referred to above.

It should be mentioned that unlike power supply conductors, grounding and EGC
are not protected by inline circuit breakers or fuses. Hence it is necessary to ensure
that their cross-sectional areas will safely carry the fault currents for their duration.

5.2.1 the area of land


The land must be suitable for placing the grounding electrodes and the resistivity
of its soil must be reasonable. If the resistivity is high, the area available is limited,
and the driving of rod electrodes to large depths is impossible due to the existence of
rocky layers, then it becomes necessary to treat the earth surrounding the electrodes
to reduce the soil resistivity.
The soil resistivity depends to a large extent on its moisture content, the amount
of salts present capable of forming ions and on its porosity. There are two methods to
reduce the resistivity of soils:

• Use of chemical salts.


• Increase the soil’s ability to retain water.

5.2.1.1 Use of Chemical Salts


The salts commonly used for improving soil resistivity are, in order of preference,

• Magnesium sulfate (MgSO4)


• Copper sulfate (CuSO4)
• Calcium chloride (CaCl2)
• Sodium chloride—common salt (NaCl)
• Potassium nitrate (KNO3)

Magnesium sulfate is the most commonly used salt because it is cheap, has a high
conductivity, and a low corrosiveness. Common salt is also cheap and can be used
if there is no danger of corrosion. Potassium nitrate and calcium chloride pose seri-
ous health and safety risks and are condemned as hazardous substances by OSHA.
Copper sulfate is extremely corrosive to steel, iron, and galvanized pipes, and is
considered environmentally unfriendly.
Two common methods used to add chemicals for the improvement of soil resistiv-
ity are shown diagrammatically in Figures 5.3 and 5.4. In Figure 5.3 the salt is placed
in a trench surrounding the electrode after which the trench is covered with a layer
of soil. As an alternative to a trench, a tile pipe of about 20 cm diameter and 60 cm
length is buried in the ground surrounding the electrode as shown in Figure 5.4. One
half of the pipe is then filled with magnesium sulfate and water is added to fill the
pipe. The pipe is provided with removable cover and should be checked periodically
and refilled if necessary.
48 An Introduction to Safety Grounding

30 cm

50 cm
30 cm

grounding conductor
welded to rod
salt placed in trench
and covered with soil ground rod ≥ 3m

FIGURE 5.3 Salt trench surrounding a ground rod.

20 cm
removable cover

tile pipe
60 cm

salt solution

ground wire rod electrode

FIGURE 5.4 Treatment with salt solution using a tile pipe.

A more recent method of using chemicals to improve soil resistivity is by using


a so-called chemical ground rod electrode (Figure 5.5). This consists of a hollow
copper tube of about 6 cm (2.5″) diameter with an exothermically welded pigtail
for connection to the grounding conductor. The tube is filled with an electrolytic
salt mixture (supplied by the manufacturer); small holes distributed along the tube
length allow the hygroscopic salts to absorb water from the surrounding soil and
Grounding Systems 49

cover

removable
cover
soil
copper tube

exothermic weld seep hole


pigtail connexion

electrolytic
back fill salts

FIGURE 5.5 Chemical grounding rod.

atmosphere. This causes the salts to dissolve and the resulting electrolytic solution
seeps out of the hole at the bottom of the tube into the surrounding soil. To further
reduce the resistivity the electrode is surrounded by a backfill marketed commer-
cially as a ground enhancement material (GEM) or ground augmentation fill (GAF);
these are noncorrosive carbon-based materials that improve grounding effectiveness.
It is claimed that a resistivity as low as 0.12Ω·m can be achieved with such installa-
tions. Both vertical and horizontal (L-shaped) rods are commercially available; the
latter are used at sites where the ground is rocky or for internal installations.
It is apparent that the use of chemical salts for improving the soil resistivity is a
temporary measure, since rain and natural drainage will gradually leach away the
salt which must be replaced periodically. The frequency of replenishment obviously
depends on the amount of rainfall and on the porosity of the soil; on the average it is
about every 2 years with the intervals becoming longer as the soil becomes more and
more conductive. If follow-up and maintenance is slack, as it is in certain countries,
it is advisable not to use this method however economical it might be.

5.2.1.2 Increasing the Soil’s Ability to Retain Water


Heavy rainfall and frequent watering lead eventually to the depletion of salts from
the soil and a sharp rise in its resistivity. To improve the soil’s capability of retaining
water, bentonite is added. Bentonite is a kind of soft porous volcanic clay which has
the property of absorbing water from its surroundings and retaining it for a consider-
able time. When placed around earth electrodes, as well as around the conductors
joining them together, it reduces the soil resistivity and hence the electrode resistance
to ground.
Bentonite is available as a dry powder which is placed directly around the elec-
trodes and the associated conductors. It is then saturated with water and covered with
about 30 cm of ordinary soil. One of the disadvantages of bentonite is that when it
is saturated with water its volume increases several times, so that care must be exer-
cised when using it in paved areas or near buildings where freedom of expansion is
limited; the resulting stresses may lead to dangerous cracks. The other disadvantage
is that when bentonite dries its volume shrinks considerably leading to a complete
50 An Introduction to Safety Grounding

loss of intimate contact between it and the electrodes, thereby prejudicing the entire
grounding system. In order to overcome these disadvantages it is best to use a mix-
ture having the following composition:

75% gypsum (hydrous calcium sulfate)


20% bentonite
5% sodium sulfate

This mixture is available commercially, its price is reasonable, and it is preferable to


use it rather than use chemical salting.

5.2.2 ground electrodeS


5.2.2.1 Driven Rods
Since the metal used for the ground electrode does not affect its resistance to ground,
the choice of metal will depend entirely on its resistance to corrosion in the type of soil
in which it will be buried. Long practical experience as well as laboratory tests have
shown that copper is the best metal as far as resistance to corrosion is concerned. The
driven rod is the most common type of electrode used for grounding and it is made
either of solid copper or of copper-clad steel. The latter is made by molecularly bond-
ing 99.9% pure copper onto a low carbon steel core and is commercially known as a
copperweld or copperbond rod. The steel provides the required mechanical strength
which enables the rod to be driven by power hammers to great depths without dam-
age, while the copper cladding protects the rod from electrolytic corrosion and allows
the use of a copper/copper junction between it and the grounding conductor. Stainless
steel rods are used in applications where soil conditions are very aggressive,such as
soils with high salt content, and where galvanic corrosion (see Section 5.3) can take
place between dissimilar metals buried in close proximity. Galvanized steel rods are
also sometimes used if corrosion considerations permit their use.
Since it is true that no part of an electrical installation is more neglected or forgot-
ten than the grounding system, this part must be designed with great care using only
the best materials proven to give the most reliable service for the longest period of
time. Copper is such a material and should therefore always be the first choice for
earth connections and for electrodes (solid copper or copper-clad).
As the resistance to ground of a rod electrode depends essentially on its length, it
has been agreed that the rod length should be equal to or be a multiple of a standard
length. This standard length is 8 ft (2.4 m) in the British system and 3 m in the metric
system. To drive an electrode to large depths two or more standard lengths are con-
nected together by means of a special coupling. Since the diameter of the rod has no
significant influence on its resistance to ground (doubling the diameter will lower
the resistance value by only 9.5%) its value is determined by the mechanical rigidity
required for driving it into the ground. In general the rod diameter should not be less
than 12.5 mm (½˝); the size most commonly used is 16 mm (5/8˝).
The most common method of driving the rods into the ground is by using a
mechanical hammer especially if the number of rods is large or the soil is not sandy
Grounding Systems 51

or the length of the rod is more than 3 m. Otherwise a hand mallet can be used to
drive the rods.
In most cases it will be found that more than one rod will be needed to obtain
the desired resistance to ground. In such cases the distance between the electrodes
should not be less than their driven depth. In small transformer substations four or
more grounding rods are driven around the perimeter of the installation in the form
of a square or rectangle; whenever possible it is preferable to carry this out at the bot-
tom of the site excavations.
In installations with four or more rod electrodes all the rods must be connected
together in a closed circuit by means of grounding conductors. The circuit is then
connected to the ground connection provisions by means of two or more conductors
in parallel.
Figure 5.6 gives typical examples of ground rod connections.

driving
stud inspection pit

coupling

rod

coupling

FIGURE 5.6  Constituents of a driven rod and typical installations [Furse].


52 An Introduction to Safety Grounding

5.2.2.2 Concrete-Encased Electrodes


The resistivity of concrete below ground level is about 30 Ω·m at 20°C. This is
lower than the average value of soil resistivity which is 100 Ω·m. Consequently in
earth of average or higher resistivity, the encasement of a rod electrode in concrete
(Figure 5.7) gives a lower resistance to ground than a similar electrode buried directly
in the ground. This is because the lowering of the resistivity of the material closest
to the electrode has the same effect as the chemical treatment of the soil around the
electrode (Section 5.2.1).
It should be mentioned here that there are commercial products which when
added to cement in place of sand and aggregate produce a conductive concrete whose
resistivity is extremely low. One such product is Marconite1 which is mixed, instead
of sand, with ordinary Portland cement in the ratio of 2:1 and then with water to form
a fairly dry mix. When used with a copper rod the mix will adhere to the rod and set
into a permanent hardened form whose resistivity is as low as 0.1 Ω·m.
The widespread use of steel reinforcing bars in concrete foundations and footings
provides a ready-made supply of grounding electrodes at structures utilizing this
construction. It is only necessary to bring out an adequate electrical connection from
a main reinforcing bar of each such footing to the building ground bus or structural
steel (Figure 5.8). The large number of such footings inherent to buildings will pro-
vide a net ground resistance considerably lower than that provided by made electrode
methods. Concrete-encased steel rods placed in excavations in rock or very rocky
soil have been found to be superior to other types of made electrodes. Such types of
electrodes constitute the ground electrodes for the majority of steel towers of high-
voltage transmission lines.

topsoil
ground rod

marconite different
concrete soil
layers

FIGURE 5.7 Ground rod encased in concrete.

1 Marconite is a registered trade mark of Marconi Communications Ltd., and is manufactured by Carbon
Int. Ltd.- Conductive Products Division.
Grounding Systems 53

grounding conductor
steel reinforcing bar
not less than 13 mm
diameter and at least nonmetallic
6 m long. protective sleeve
approved
connection

≥ 50 mm
foundation in direct
contact with earth

FIGURE 5.8 Concrete-encased bar [NEC 250.52].

5.2.2.3 Water and Gas Pipes


It is possible to utilize water pipes as auxiliary earthing electrodes provided that the
following conditions are met:


1. The pipes are metallic

2. The pipes are buried in the ground and the buried length exceeds 3 m.

3. There is electrical continuity. If there are junctions or meters they must be
provided with suitably sized jumpers as shown in Figure 5.9.

through the ground bus-bar.

Water pipes should never be used as the sole means of grounding because of the pos-
sibility of replacement of some sections by plastic pipes and of non-strict adherence
to electrical continuity.

braided copper strap

metal pipe

approved clamp jumper

weld to ground connection


provisions

jumper > 3m
jumper

coupling between
> 3m pipe ends

FIGURE 5.9 Jumpers on metal pipes to ensure continuity.


54 An Introduction to Safety Grounding

Pipes carrying gas or any flammable liquids should never be used as part of a
grounding system through which current flows.

5.2.2.4 Structural and Reinforcing Steel


The metal frame of a building or structure as well as steel reinforcing bars bonded
together can be used as a ground electrode. However, the resistance to ground must
be measured to make sure that additional electrodes are not required.

5.2.2.5 Equipotential Bonding


To equalize the potential to which all grounded bodies rise, all means of grounding
at a given site must be bonded together to form an equipotential grounding electrode
system as shown in Figure 5.10.

5.2.3 connectionS in grounding SyStemS


Joints and connections are critical elements in the grounding circuit through which
the short circuit current flows. It is therefore of the utmost importance that they be
made with the utmost skill to ensure the integral safety of the grounding system.
There are three methods of connecting the grounding conductors to the ground
electrodes, and connecting the EGC to the metal structure of the equipment or to the
ground connection provisions. These methods are as follows:

1. Mechanical connection using clamps and bolts.

Both clamp and bolt should be made of the same metal as the electrode and conductor
to prevent galvanic corrosion and the connections must be protected against any acci-
dental damage and designed for easy inspection. The most important requirement for

down conductor

ground bus

water and gas meters to lightning ground

to protective ground

telephone cable
power cable
water
gas

FIGURE 5.10 Equipotential ground system.


Grounding Systems 55

clamp connections is the establishment of intimate and permanent contact between


the metal surfaces. Therefore, when making a connection the surfaces must be
smooth, clean and free from any insulating layer, and the tightening bolts must exert
sufficient pressure to maintain good contact in the presence of any stresses or shocks
or vibrations associated with the equipment or the surrounding environment. If it is
necessary to join two different metals then all necessary precautions (recommended
by the manufacturer) must be taken to protect the connection against galvanic corro-
sion. Figure 5.11 gives some commonly used types of clamps.

2. Welded connection.

This is one of the best methods to make permanent connections and has the advantage
that conductors having a smaller cross sectional area can be used (see Section 5.4.2.1).
There are several methods used for welding: brazing, silver soldering, and exothermic
welding also known as thermit welding. Soldering should never be used. All types of
welding, except exothermic welding, require experience and high skill.
The best method of welding is the exothermic one and it has the advantage of
welding copper to steel or iron. In this method a powder consisting of a mixture
of aluminum and copper oxide is placed around the joint inside a graphite mould.
When the mixture is ignited by a starting powder, the temperature generated by the
exothermic reaction reaches more than 2,000°C and the copper oxide is reduced to
molten copper which flows through the mold onto the conductors to be joined. One
of the advantages of this method is that it can be used to weld conductors to ground

conductor

lock
washers
metal sheet

ground rod clamps

clamps for connecting


clamps for connecting conductor conductors with different
to metal structure e.g. tower cross sections

FIGURE 5.11 Some common types of clamps [Furse].


56 An Introduction to Safety Grounding

cable to electrode

cable to rectangular bar

stud to steel surface

cable to cable

cable to steel surface or pipe

FIGURE 5.12 Welded connections [Furse].

electrodes in places where it would be difficult to use the other methods of welding;
another advantage is that it does not require a skilled welder. However, because a dif-
ferent mold is needed for each different joint configuration, this method is economic
only if there are a large number of joints which are similar. Figure 5.12 shows a
number of welded joint configurations.

A special copper or copper-alloy sleeve is compressed simultaneously onto the


ground rod and the grounding conductor by means of a special hydraulic press. This
is the most economical method of connection and has all the advantages of thermit
welding.
Whichever method is used, one must always ensure that all surfaces are clean and
free from grease, paint, or any other insulating film.

5.2.4 grounding conductorS


Grounding conductors usually consist of buried copper cables. If bare cables are used
they may cause galvanic corrosion of other metals buried in the ground. However, for
short cable lengths buried near the surface in dry soil of high resistivity, the corrosion
Grounding Systems 57

aspect may be neglected. Long lengths of copper cable buried in relatively moist soil
of low resistivity should have a waterproof cover.
Aluminum, or any other highly anodic metal, should not in any circumstances be
used for grounding conductors.

5.2.4.1 Size of Grounding Conductors


Since grounding conductors do not carry current except under short circuit condi-
tions, the choice of the cross-sectional area of the conductor is determined by the
magnitude of the fault current and its duration (i.e., the speed with which the pro-
tective device disconnects the fault) as well as by mechanical strength if exposed
to possible damage. If unprotected then the minimum size to be used is 16 mm2 in
noncorrosive environments and 25 mm2 in corrosive environments.
It is possible to specify the cross-sectional area of grounding conductors accord-
ing to either European or US wire sizes. In European specifications areas are given
in mm2 or in.2 (for British specifications) whereas in US specifications areas are given
in circular mils. A circular mil is the area of a circle whose diameter is 0.001 in., i.e.,

1cmil = 7.854 × 10 –7 in.2 = 5.07 × 10 –4 mm 2

1mm 2 = 1973.5 cmil

Because a circular mil is such a small unit, the cross-sectional areas of conductors
are usually given in MCM units (1 MCM = 1 kcmil = 103cmil). For the purpose of
comparison, Tables A.1 and A.2 in the Appendix give the cross-sectional areas and
diameters of conductors according to the different specifications.
The cross-sectional area of a copper grounding conductor must not be less than the
value determined from the following equation (known as the Onderdonk equation)
which is based on the assumption that the heating process is adiabatic, i.e., there is
no heat lost to the surrounding medium, an assumption which is valid provided that
the duration of the short circuit current does not exceed 5 s:

 T − Ta 
log10  m + 1
 234 + Ta 
I = A′
33t

A ′ = I t C cmil

A′ I t
A= = mm 2 (5.1)
1,973.5 k

where
A΄ = conductor area in cmil,
A = conductor area in mm2,
I = short circuit current (A),
58 An Introduction to Safety Grounding

t = current duration (s),


Ta = ambient temperature (assumed 40°C),
Tm = maximum allowable temperature (°C) which is determined by the type of
joints,
k, C = constants whose values depends on type of joint. For copper conductors
their values are as follows:

C k
11.6 170 clamp joints with threaded bolts (Tm = 250°C)
9.11 217 brazed joints (Tm = 450°C)
6.96 284 compression or exothermic soldered joints (Tm = 1,083°C)

The melting point of copper is 1083oC.


The dependence of the conductor cross section on the current duration indicates
that the type of protective devices used and their current/time interruption char-
acteristics must be taken into consideration in designing protective earthing (PE)
systems.

Example:

Assume that the magnitude of the short circuit current is 20,000 A and that the
protective device clears the fault after five cycles, i.e., 0.1 s.

20, 000 × 0.1


A= = 37.2 mm 2
170
A′ = 20, 000 × 0.1 × 11.6 = 73, 365 cmil

From Tables A1 and A2 we find that


Amin = 50 mm2
′ = No. 1 AWG (American Wire Gauge)
Amin

A = 29.14 mm2
A' = 57,617 cmil
From Tables A1 and A2 we find that
Amin = 35 mm2
′ = No. 2 AWG
Amin

A = 22.7mm2
A' = 44,019 cmil
Grounding Systems 59

From Tables A1 and A2 we find that

Amin = 25 mm2
′ = No.3 AWG
Amin

It is apparent from this example that compression and exothermic welding give
the best type of joint since the conductor cross section for such joints is 50% less
than that for mechanical joints and 30% less than that for brazed joints.

5.2.4.2 Protective Sleeve


In some installations it is necessary to protect parts of the grounding conductor path
from mechanical damage to which the conductor could be exposed. In these cases
the conductor is placed inside a protective sleeve which has sometimes to be a metal
one. For these types the sleeve has to be connected to the conductor at both ends as
shown in Figure 5.13. The reason for this is that under short-circuit conditions the
large current flowing in the conductor produces a strong magnetic field in the sleeve,
especially if it is made from a magnetic material such as steel. This field will generate
in the sleeve eddy currents which heat it to temperatures that may attain its melting
point, causing damage to the conductor. Moreover, the presence of a magnetic sleeve
around the conductor causes a large increase in its self-inductance and hence in the
impedance of the grounding conductor. To avoid all such effects both ends of the
sleeve must be connected to the conductor with wire of the same size as the conduc-
tor itself. The majority of metal sleeves consist of steel pipes and since the larger
part of the current passes through the pipe (because of the skin effect) its electrical
resistance must not be less than that of a similar length of grounding conductor.
Table 5.1 shows the approximate equivalent sizes of steel pipe and copper conductor.

welded wires

conductor
metal sleeve

FIGURE 5.13 Protective metal sleeve with jumpers at both ends.

TABLE 5.1
Equivalent Sizes of Steel Pipe and Copper
Conductor
Copper conductor
Steel pipe (inch) mm 2 AWG
½ 25 No. 4
¾ 35 No. 2
1 50 No. 1
1¼ 70 No. 2/0
60 An Introduction to Safety Grounding

5.2.5 ground connection proviSionS


Except for very small installations, it is not feasible to connect the grounding
conductors directly from the ground electrodes to the equipment. The normal
procedure is to terminate the buried grounding conductor at ground connection
provisions. These provisions consist of plates or bus bars located at readily accessible
locations in the installation. Ground plates are used in concrete structures such
as equipment foundations, vaults, manholes, piers, etc. The plates are placed at a
suitable site on the concrete structure such that the flat surface is flush with the con-
crete surface. Grounding conductors are connected to the plates by means of spe-
cial clamps to which the conductors are either welded or fixed with bolts. If the
installation of ground plates is not practical then copper bus-bars are used. These
bars are fixed to the surface of walls at convenient locations for either permanent
or temporary ground connections. They must not be fixed to any metal surface and
the diameter of the fixing hole on the bar itself must not exceed one quarter of the
bar width. The size of the bus bars is determined as for the grounding conductors;
for distribution stations at which the short circuit current does not exceed 22 kA, the
minimum cross section of the ground bus is 125 mm2 (5 × 25 mm). Connections to
the bus bar may be either mechanical or welded.
In installations where fuses, disconnect switches, cable trays, etc., are mounted
on metal beams, these beams may be used as ground connection provisions for such
equipment provided that the beams are properly grounded. They may also be used for
the temporary grounding of minor equipment not mounted on the beams. However,
major pieces of equipment, such as transformers and circuit breakers, have to be
grounded through proper ground connection provisions even if they are mounted on
grounded beams.
It should be mentioned here that the standard specifications of many countries
do not permit the passage of current except in paths confined to a specific electric
circuit.

5.2.6 equipment grounding conductorS


These are the conductors which extend between the metal parts of the equipment to
be grounded and the ground connection provisions referred to above. Several means
are used to ground equipment:

1. Bare or covered copper cable


2. Rigid metal conduit or raceway
3. Flexible metal conduit (FMC)
4. Metal cable trays
5. Shield and armor of some cables

The most common means is copper cable.


The cross-sectional area of EGC used in factories and stations is determined by
using Eq.(5.1) exactly as for grounding conductors.
Grounding Systems 61

In residential and commercial buildings EGC are the protective conductors, which
are usually identified by an insulating cover which is green or green with one or more
yellow stripes. The method for determining the size of these conductors depends
upon which national specification is used. In what follows we shall give three of the
most commonly used methods.

5.2.6.1 Conductor Size according to the Adiabatic Equation


The conductor cross section is determined from the following (adiabatic) equation:

I t
A= mm 2 (5.2)
k
where
A = conductor area in mm2,
I = short-circuit current (A),
t = current duration (s), which is the disconnection time of the protective device,
k = a factor which depends on the material of the conductor, the insulation and the
initial, and final temperatures.

Equation (5.2) is often expressed as

aI n t
A= mm 2 (5.2a)
k
where
In = the rated current of the protective device and
aIn = multiples of the rated current.

Equation (5.2) is known as the adiabatic equation because it is based on the


assumption that for the duration of the short-circuit current there is no loss (or gain)
of heat. This means that the entire ohmic losses generated during the flow of this cur-
rent are used to heat the conductor without any loss to the surrounding medium. This
assumption is valid provided that the duration of the short circuit is less than 5s and
since all modern LV circuit breakers and fuses have short circuit tripping times very
much shorter than that (a few hundred milliseconds), using the adiabatic equation
greatly simplifies calculations. The equation is derived as follows.
For a short-circuit current I with a duration time t the energy dissipated by the
ohmic losses is

Er = I 2 Rt

This energy will be converted to heat raising the conductor temperature by ∆θ.
This thermal energy is given by

Eth = mass × specific heat × temperature rise



= mc ∆θ
62 An Introduction to Safety Grounding

where
m = mass of cable conductor (gm)
c = specific heat of conductor material (J/gm·°C)
∆θ = maximum permissible temperature rise (°C)

Since the process is assumed to be adiabatic the two energies Er and Eth must be
equal so that
t θ2



0

I 2 Rdt = mc dθ
θ1
(5.3)

and assuming that the resistance is constant

I 2 Rt = mc ∆θ

where

R = ρr l / A; m = ρd A l

and
ρr = resistivity of conductor (Ω·mm)
ρd = density of conductor material (gm/mm3)
A = cross-section area of conductor (mm2)
Substituting these values in the above equation and solving for A

I t
A=
k
where

Q ∆θ
k= (5.4)
ρr

and

Q = c ρd = volumetric heat capacity (J/°C mm3)

For convenience gm and mm units are used to give the cross-sectional area A in mm2.
Table 5.2 gives the values of the relevant properties of conductor materials.
These values may vary very slightly for copper and aluminum depending on the
manufacturing process. For steel, however, these values may vary considerably,
resistivity 104–186 µΩ∙mm, specific heat 410–500 J/°C. The values given are for high
conductivity steel wire used as armor for cables and also serving as PE conductor.
Q may be considered constant since the increase in specific heat with increasing
temperature is compensated by an almost equivalent decrease in density.
In order to determine the value of the k factor in Eq.(5.4) it is necessary to
determine what value to use for the resistivity ρr which is not constant but increases
Grounding Systems 63

TABLE 5.2
Conductor Parameters
Coefficient Coefficient
Volumetric of of
Specific capacity Resistivity Resistivity Resistivity
Conductor Density heat Q ρ20 α20 α0
Material (gm/mm3) (J/gm°C) (J/°C mm3) (μΩ mm) (/oC) (/oC)
Copper 8.96 × 10–3 0.3853 3.45 × 10–3 17.24 × 10–6 3.93 × 10–3 4.26 × 10–3
Aluminum 2.70 × 10–3 0.921 2.50 × 10–3 28.26 × 10–6 4.03 × 10–3 4.38 × 10–3
Steel 7.86 × 10–3 480 3.8 × 10–3 138 × 10–6 4.50 × 10–3 4.95 × 10–3

with increase in temperature. The variation of resistivity with temperature is


approximately given by

ρθ = ρ0 [1 + α 0θ ] = α 0 ρ0 ( β + θ ) (5.5)

where α 0 is the temperature coefficient of resistivity at 0°C and β = 1/α 0. Using this
value of ρ θ Eq. (5.3) now becomes

θm
A2 Q dθ
∫ I dt = I t =
2 2
ρ0 ∫ 1+α θ
θi
0

A 2 Q 1 + α 0θ m
= ln
ρ 0 α 0 1 + α 0θ i
= A2 k 2 (5.6)

where θi is the initial temperature and θm is the maximum temperature the insulation
can withstand without damage. Substituting for α 0 ρ 0 from Eq. (5.5) and assuming
that θ = 20°C we obtain finally

Q(β + 20) (β + θ m )
k= ln (5.7)
ρ20 (β + θ i )

This equation is used by a number of standards and the values of k shown in Table 5.3
are derived from it. Table 5.4 gives the values of k for bare conductors also derived
from it as given in the standard IEC60364-5-54.
An alternative method (proposed by the author) of determining the factor k from
Eq. (5.4) is to assume a resistivity whose value is the average of its values at the
temperature limits imposed by cable type and installation conditions as obtained
from Eq. (5.5). It is readily shown that this average is given by

ρav = 1 2 ρ20  2 + α 20 (θ i + θ f – 40 )  (5.8)


64 An Introduction to Safety Grounding

TABLE 5.3
Values of k Derived from Eq.(5.6)
Insulation Material and Initial and Final Temperature
(°C)
Conductor Installation XLPE Butyl
Material Conditions PVC EPR Rubber
30°C 160°C 30°C 250°C 30°C 220°C
Cu Individual insulated 143 176 166
Al conductor or bare 95 116 110
Steel conductor in contact 52 64 60
with adjacent cable
covering
70°C 160°C 90°C 250°C 85°C 220°C
Cu Conductor within a 115 143 134
Al multicore cable 76 94 89
60°C 170°C 80°C 200°C 75°C 220°C
Cu Sheath or armor used as 141 128 140
Al grounding conductor 93 85 93
Steel 51 46 51

PVC, polyvinylchloride (70°C)—thermosetting; XLPE, cross-linked polyethylene—thermosetting; EPR,


ethylene-propylene rubber—thermosetting; and butyl rubber (85°C)—thermosetting.

TABLE 5.4
Values of k for Bare Conductors Where There Is No Risk of Damage to Any
Neighboring Material by the Temperatures Indicated
Conditions
Material Visible and in
of Conductor Restricted Areasa Normal Conditions Fire Risk
Copper Temp. max 500oC 200oC 150oC
k 228 159 138
Aluminum Temp. max 300oC 200oC 500oC
k 125 105 91
Steel Temp. max 500oC 200oC 150oC
k 82 58 50

Reproduced from IEC 60364-5-54:2011 with permission.


The initial temperature of the conductor is assumed to be 30°C
a The temperatures indicated are valid only where they do not impair the quality of the connections.
Grounding Systems 65

Thus
For copper conductors for which,

Q = 3.45 × 10 –3 J / °Cmm 3 , ρ20 = 17.24 µΩmm, α 20 = 3.93 × 10 –3

with cross-linked polyethylene (XLPE) insulated individual conductors


(30°C–250°C):

ρav = 25 µΩmm, k = 174

with polyvinylchloride (PVC) insulated conductor within a cable (70°C–160°C):

ρav = 24.1 µΩmm, k = 114

For bare conductor (30°C–200°C):

ρav = 23.7 µΩmm, k = 157

For aluminum conductors for which,

Q = 2.5 × 10 –3 J / °Cmm 3 , ρ20 = 28.26 µΩmm, α 20 = 4.03 × 10 –3

with XLPE-insulated individual conductors (30°C–250°C):

ρav = 39.05 µΩmm, k = 115

with PVC-insulated conductor within a cable (70°C–160°C):

ρav = 39.05 µΩmm, k = 76

These values for k are almost identical to those derived from Eq. (5.6). We may thus
rewrite Eq. (5.4) as,

Q∆θ
k= (5.9)
ρav
It should be pointed out here that Eq. (5.2) for EGC is similar to Eq. (5.1) for grounding
conductors except that the factor k is defined differently. They are both based on the
adiabatic assumption which is valid for short-circuit durations less than 5 s.
The duration of the ground fault current depends on the time taken by the protective
device (circuit breaker or fuse) to interrupt the current. This duration is primarily
determined by the tripping characteristics of the breakers and the time-current char-
acteristics of the fuses used and which are normally supplied by the manufacturer.
Figures 5.14 and 5.15 show typical time-current characteristics for circuit-breakers
and fuses.
66 An Introduction to Safety Grounding

overload
release

10

setting range
1 (2 − 10) In
time,s

instantaneous
0.1 release

In
aIn (multiples of rated current In)

FIGURE 5.14 Tripping characteristics of circuit breakers.

In aIn

10
time,s

0.1

10 102 103
short-circuit current, A

FIGURE 5.15 Time-current characteristics of fuses.

The majority of modern circuit breakers and fuses have fault clearing times
around 100ms thus justifying the adiabatic assumption on which Eqs. (5.2) and
(5.4) are based. During this period the DC component in the initially asymmetric
short-circuit current would have decayed so that the interrupted current may be con-
sidered symmetrical.
Short-circuit currents may range from 0.5 to 30 kA depending on the system
parameters and distance of fault from the supply source. Table 5.5 shows the
grounding conductor size per kA fault current for different values of the factor k
given in Table 5.4 and for the time range 0.05–0.5s. For nonstandard cross-sectional
areas calculated from the adiabatic equation (5.2) the nearest higher standard section
is to be chosen.
Grounding Systems 67

TABLE 5.5
Effect of Short-Circuit Duration t on Conductor Cross-Sectional
Area per kA Short-Circuit Current
Conductor Area(mm2/kA)
k 0.05 s 0.1 s 0.2 s 0.3 s 0.5 s
176 1.27 1.80 2.54 3.13 4.02
166 1.35 1.90 2.66 3.30 4.26
143 1.57 2.21 3.13 3.80 4.94
134 1.67 2.36 3.34 4.09 5.28
116 1.93 2.72 3.85 4.72 6.09
110 2.03 2.87 4.06 4.98 6.42
95 2.36 3.26 4.71 5.77 7.44
89 2.52 3.55 5.02 6.16 7.94
76 2.95 4.16 5.88 7.21 9.30
51 4.39 6.20 8.76 10.63 13.86
46 4.86 6.87 9.71 11.91 15.37

If it is not desired to calculate the minimum cross-sectional area of a protective


conductor from Eq. (5.2) then both the standards BS 7671 and IEC 60364 differenti-
ate between two cases:

Phase conductor area (mm2) PE


A ≤ 16 A
16 < A ≤ 35 16
A > 35 ½A

5.2.6.2 Conductor Size According to USA Specifications (NFPA-70)


According to the National Electric Code the cross-sectional area of the EGC is
determined by the ampere rating of the protective device (fuse or circuit breaker)
installed in the circuit feeding the equipment. Table 5.6 gives the relationship
between the size of the grounding conductor and the rated current of the protective
device. Figure 5.16 shows the sizes of the different conductors according to this table.
When there is more than one circuit in the same conduit a single EGC may be
used (Figure 5.17) and its size chosen according to the rated current of the largest of
the protective devices associated with the conductors inside the conduit.
68 An Introduction to Safety Grounding

TABLE 5.6
Minimum Size of EGC
Rated Current of Protective Device (A) Conductor Sizea(copper)
15 14 AWG
20 12
30 10
40 10
60 10
100 8
200 6
300 4
400 3
500 2
600 1
800 1/0
1000 2/0
1200 3/0
1600 4/0
2000 250 MCM
2500 350
3000 400
4000 500
5000 700
6000 800

Reprinted with permission from NFPA 70-2017, National Electrical Code,


Copyright © 2017, National Fire Protection Association, Quincy, MA. This
reprinted material is not the complete and official position of the NFPA on the
referenced subject, which is represented only by the standard in its entirety
which may be obtained through the NFPA website www.nfpa.org.
a Size not required to be larger than the circuit conductors which supply the

equipment (See Tables A.1 and A.2 for relation between conductor size and
area).

5.2.6.3  Metal Cable Trays as Grounding Conductors


Cable trays are extensively used in wiring systems in preference to conduits because of
their safety features and cost savings. Metal trays are either aluminum or ­galvanized
steel or stainless steel and the ladder type cable tray (Figure 5.18) with 9″ (23cm)
between rungs is the most widely used as it provides excellent ventilation for heat
dissipation and easy exits and entrances for cables. Metal trays can be used as EGC
provided that certain conditions are met. These are given as follows:
Grounding Systems 69

feeders (protective conductor)


supply panel 1 panel 2
300 A

100 A 30 A
4 AWG
4 AWG 4 AWG
equipment grounding
conductor
load load
8 AWG 10 AWG

FIGURE 5.16 Size of EGC according to Table 5.6.

30 circuit 1
More than one circuit
conductor inside same 100 circuit 2
conduit with common
grounding conductor 40 circuit 3

8 AWG
common grounding conductor

FIGURE 5.17 Several circuit conductors and common protective conductor inside same
conduit.

jumper to ground terminal

to panel
board

tray cross section

FIGURE 5.18 Cable tray as grounding conductor.


70 An Introduction to Safety Grounding

1. Manufacturer must provide a label showing cross-sectional area available.


2. The minimum cross-sectional area of the trays shall conform to the NFPA
requirements given in Table 5.7. This area depends on the maximum rating
of the protective device (fuse, circuit breaker, or protective relay trip setting)
for any cable circuit in the tray system.
3. Electrical continuity is of primary importance so that any mechanically
­discontinuous sections have to be bonded either by jumpers with size
­according to grounding conductor specifications, or by using special tray
connectors.

If the tray cannot be used as EGC then an independent conductor is to be used as


EGC either as a separate conductor or a conductor within a multicore cable. Since
the tray must be grounded anyway, the EGC should be connected to the tray at the
beginning and at the end and every 50ft (15m); the tray can then act as an auxiliary
conductor in parallel with the main EGC.
Connections between metal trays and any copper ground provisions must be made
using special strips to avoid galvanic corrosion (Section 5.3).

TABLE 5.7
Metal Area Requirements for Cable Trays Used as EGC
Maximum fuse ampere rating, circuit breaker ampere Minimum cross-sectional area of metala
trip setting, or circuit breaker protective relay ampere Steel tray Aluminum tray
trip setting for ground fault protection of any cable
in2 mm2 in2 mm2
circuit in the cable tray system.
60 0.2 1.3 0.2 1.3
100 0.4 2.6 0.2 1.3
200 0.7 4.5 0.2 1.3
400 1.0 6.45 0.4 2.6
600 1.5 9.68b 0.4 2.6
1000 − − 0.6 3.9
1200 − − 1.0 6.45
1600 − − 1.5 9.68
2000 − − 1.5 13b

Reprinted with permission from NFPA 70-2017, National Electrical Code, Copyright © 2017, National
Fire Protection Association, Quincy, MA. This reprinted material is not the complete and official position
of the NFPA on the referenced subject, which is represented only by the standard in its entirety which may
be obtained through the NFPA website www.nfpa.org.
a Total cross-sectional area of both side rails of ladder or trough cable trays, or the minimum cross-­

sectional area of metal in channel cable trays or cable trays of one-piece construction.
b Steel cable trays shall not be used as EGC for circuits with ground-fault protection above 600 A.

Aluminum shall not be used as EGC for circuits with ground-fault protection above 2000 A.
Grounding Systems 71

FIGURE 5.19 Flexible metal conduit (FMC).

5.2.6.4 Flexible Metal Conduits


As the name implies FMC (Figure 5.19) is used wherever flexibility and protection
are required, as well as for reducing the number of fittings in an installation or to
minimize the transmission of vibrations from equipment such as motors. According
to the Institute of Engineering and Technology (IET) Wiring Regulations “flexible or
pliable conduit shall not be used as a protective conductor.”
However, the use of such conduit as EGC is permitted by the NEC. There are
three types of FMC:

1. FMC, which is a conduit of circular cross section made of helically wound,


formed, interlaced metal strip.
Liquid-tight flexible metal conduit—LFMC, which is conduit of circular
cross section having an outer waterproof, nonmetallic, sun-light resistant
jacket over an inner flexible metal core.
2. Flexible metallic tubing—FMT, which is conduit of circular cross section,
flexible, metallic, and liquid tight without a nonmetallic jacket.

The above types can be used as EGC if they meet all the following conditions:


• Fittings used (connectors, couplings, terminations, etc.) are listed for
grounding.
• The total length of flexible piping in the ground circuit does not exceed
1.8 m (6 ft). For greater lengths an EGC must be installed with the cir-
cuit conductors.

• The circuit conductors contained in the conduit are protected by over-
current devices rated at 20 A or less.
72 An Introduction to Safety Grounding

• If used to connect equipment whose flexibility is necessary after instal-


lation, an EGC must be installed.

• For ⅜ and ½ in. trade sizes2 the circuit conductors contained in the con-
duit are protected by overcurrent devices rated 20 A or less.
• For trade sizes ¾ through 1¼ in. the circuit conductors contained in the
conduit are protected by overcurrent devices rated not more than 60 A and
the grounding path does not contain any flexible piping of smaller size.
• If used to connect equipment whose flexibility is necessary after instal-
lation, an EGC must be installed.

5.2.6.5 Cable Sheath and Armor


The sheath or armor of cables can be used as protective grounding conductor in
which case the manufacturer’s data must be consulted to determine the sheath and
armour electrical specifications (see also Section 5.5.3). It should be mentioned that
this also applies to metal raceways used to enclose cables and wires.

5.2.6.6 Jumpers
Jumpers are used to ensure the electrical continuity of the connection to ground
of both grounding conductors and EGC as well as to ensure continuity of water
pipes used as auxiliary grounding electrodes (see Figure 5.9). Their size is deter-
mined in a manner similar to that for the size of grounding conductors or EGC
described above.

5.3 METALLIC CORROSION


If two different metals are present in a moist medium or if two different metals are
in contact in a moist environment, with the passage of time one of the metals will
corrode. The reason for this is the electrochemical process which results in the corro-
sion of the more anodic or less noble of the two metals. Table 5.8 gives the so-called
Galvanic series of some metals and alloys. The Galvanic classification of metals is
relative since the measured electrode potentials depend on the electrolytic medium
used, but the differences are not large. Table 5.8 also gives the so-called anodic index
of the different metals referred to gold as base and for sea water as electrolyte.
A metal is considered to be more anodic than another if it falls above it in the table.
For example galvanized steel is more anodic than copper (the voltage difference
between them is 0.8 V), but copper is more anodic than gold (voltage difference
0.4 V). If a galvanized steel pipe is buried near a grounding copper electrode, the
pipe will corrode but not the electrode.
The rate of corrosion depends on the galvanic potential difference and on the
moisture and salt content of the soil or the atmosphere. In coastal areas and polluted

2 Because the measured size is not the same as the nominal size, ‘trade’ sizes are used rather than
dimensions (e.g., a trade size ½ FMC has an actual inside diameter of 0.635˝).
Grounding Systems 73

TABLE 5.8
Galvanic Series for Metals and Alloys
Metal Anodic Index
Anode
Magnesium and its alloys 1.75
Galvanized steel 1.2
Galvanized iron 1.2
Aluminum 0.9
Cast iron 0.85
Duralumin 0.75
Lead 0.7
Tin 0.65
Tin-Lead solders 0.65
Chromium-plated steel (0.005˝) 0.65
Chrome steel 18/2 0.5
Copper and its alloys 0.4
Silver solder 0.35
Nickel-plated steel 0.3
Titanium 0.15
Silver 0.15
Carbon 0.05
Gold 0
Platinum 0
Cathode

industrial areas the corrosion rate is high. For any two different metals the rate of
corrosion of the more anodic of the two is directly proportional to the area of the
cathode and inversely proportional to the area of the anode. Hence, to minimize
corrosion rate the area of the more anodic metal should not be less than that of the
metal cathode (Figure 5.20).
If the ground connections consist of two different metals then the following points
must be taken into consideration:

1. The more anodic metal must not be the body of the equipment or of
the structure. For example if it is required to connect a galvanized steel
structure to copper grounding electrodes, then this must be done by means
of a galvanized steel strip which can be easily replaced if it is corroded by
galvanic action.
2. Junctions must be above the earth’s surface.
3. Junctions must be protected against moisture.
4. Junctions must be located at readily accessible sites for inspection purposes.

74 An Introduction to Safety Grounding

high current density ⇒ high corrosion rate

moist medium bad


anode

cathode

direction of electron flow

moist medium acceptable


cathode
anode

FIGURE 5.20 Effect of anode and cathode areas on corrosion rate.

5.4 GENERAL RULES FOR THE GROUNDING


OF BRANCH SUBSTATIONS
Figure 5.21 shows the different grounding systems for a simple network consisting
of a branch substation and a distribution substation feeding a low voltage (≤1,000V)
network. The different grounding systems are as follows:

A. High-voltage protective grounding system which is the grounding of all


metal parts (equipment casings, cable sheaths, etc.) which belong to the
high-voltage network.
B. Grounding of the neutral point.
C. Low-voltage protective grounding which is the grounding of all metal parts
which belong to the low-voltage network.

The question which now arises is this: should each grounding system be separate
from the others, or should a single common grounding system be used? Although

branch substation distribution station


low voltage
feeder

overhead line
or cable

B A A B C

FIGURE 5.21 Different grounding systems at transformer stations.


Grounding Systems 75

there is no clear-cut answer to this question, long experience in many parts of the
world has led to the adoption of the following recommendations.

It is recommended to use a common ground for the neutral point and for the protective
grounding. The reason for this is as follows.
If a fault occurs inside the station between a conductor and a grounded metal
body, then if the two grounding systems are separate (Figure 5.22a) the whole short
circuit current will flow into the protective ground. However, if the two ground points
do not fall outside each other’s resistance area there is bound to be a resistance cou-
pling between them as shown in Figure 5.22b. In order to clarify what the resistances
R1, R2, and Rm represent let us assume for simplicity that the earth electrodes are
represented by equivalent hemispheres of radii a and b, respectively. From Eq. (2.1)
the resistance from electrode A to electrode B is given by

ρ 1 1
R1 =  − 
2π  a D 
= Ra − Rm

A B

R1 R2

(a) (b)
I
Rm
I
A B
D

(c)

FIGURE 5.22 (a) and (b) separate grounds; (c) common ground.
76 An Introduction to Safety Grounding

where Ra = ρ/2πa is the absolute resistance of electrode A and Rm = ρ/2πD is the


resistance from absolute zero (infinity) up to a distance D from A. Similarly the resis-
tance from electrode B to electrode A is given by

ρ 1 1
R2 =  − 
2π  b D 
= Rb − Rm

where Rb = ρ/2πb is the absolute resistance of electrode A. Thus,

Ra = R1 + Rm and Rb = R2 + Rm

It is evident that when D is large compared to a and b then the mutual resistance Rm
is negligibly small.
The potential of the transformer tank T and any metal bodies connected to the
same ground will rise to a value given by I(R1 + Rm), whereas the potential of the
neutral point will rise to IRm. In this case, therefore, it is necessary to either

It should also be pointed out here that the flow of large currents between two earth
points in a limited area can lead to dangerous step voltages.
With a common ground point for both neutral and protective grounding
(Figure 5.22c) the only danger is the rise in the potential of metal parts when a short
circuit occurs. This danger may be avoided by having a grounding system with a
low resistance and fast-acting protective devices such that the potential rise does not
exceed the maximum permissible touch voltage (see Section 4.2).

II. The protective fence

The answer to the question whether the protective fence surrounding the station
should be connected to a separate ground or to the station ground depends primarily
on the relative danger to which persons or animals outside the fence are exposed if
they are touching the fence when a fault occurs, and to which persons inside the fence
are exposed if they are simultaneously touching the fence and any equipment con-
nected to the station ground when a fault occurs. If the possibility of the latter occur-
rence does not exist or if it can be prevented by some means, then it is preferable
that the fence ground should be separate from the station ground. For more details
on the practices of grounding substation fences the reader should consult IEEE Std
80-2000.
Grounding Systems 77

The majority of such stations are distribution stations. The decision whether to
have separate grounds or a common ground or two common and one separate
ground is determined primarily by the requirement that in the event of a fault to
earth on the high-voltage side, the potential of the neutral conductor does not rise
to a value which may harm people or cause fires. Since distribution substations can
have different types of installation arrangements, we shall give in what follows
the practice followed for grounding (common or separated) the various types of
installations.

(ii) When the substation is enclosed in a steel or reinforced concrete


structure, it is usually difficult to separate the low voltage from the
high-voltage grounding systems. In such cases a common ground can
be used for the two systems as shown in Figures 5.25 and 5.26.

transformer

High voltage distribution panel Low voltage

cable with
metal sheath
to low voltage
network by
overhead line or
cable
A,B,C

FIGURE 5.23 Input or output feeders cables with metal sheath.


78 An Introduction to Safety Grounding

overhead line

PME
earthing

insulating cover
A B,C
˃ 20m

FIGURE 5.24 PME of the distribution system.

metal enclosure

B
A,C insulated cable sheath
˃20m

FIGURE 5.25 Common point (A, C) for protective grounding; separate point (B) for neutral
grounding.

metal enclosure

thermoplastic insulation

A,C B
˃20m

FIGURE 5.26 Common point (A, C) for protective grounding; separate point (B) for neutral
grounding.
Grounding Systems 79

5.5 PROTECTIVE GROUNDING OF CONSUMER


INSTALLATIONS: TYPES OF GROUNDING SYSTEMS
To protect the consumer from electric shock when a fault occurs between a live
conductor and a metal body exposed to touch, the following basic requirements must
be met:

According to international standards (IEC 60479-1) the continuous permissible touch


voltages are as follows:3

For general applications,


≤50 V for alternating current
≤120 V for ripple-free4 direct current
In agricultural areas and for medical purposes,
≤25 V for alternating current
≤60 V for ripple-free direct current

The admissible limits of touch voltage as a function of time are summarized in


Table 5.9.
The types of grounding systems are internationally designated by letters. The
significance of these letters is as follows:

First letter: grounding arrangement at the supply source,


T: source has an earth (terra) point.
I: source is totally isolated or the neutral point is earthed through a high impedance.
Second letter: relationship of exposed metal enclosures and structures to earth,
T: the exposed conductive parts are directly connected to earth irrespective of the
grounding arrangement of the supply.
N: direct electrical connection of the exposed conductive parts to the grounding
point of the supply, which is usually the neutral point.

3 Any circuit in which the potential of a conductor to ground does not exceed these voltages is defined
as an extra-low voltage—ELV circuit There are essentially three types of ELV circuits: (a) separated
ELV (SELV), which is an ungrounded circuit supplied from an isolation transformer, battery, or diesel
generator and physically separated from all other circuits; SELV circuitry must be used in swimming
pools (see IEC 60364-7-702). (b) Protected ELV (PELV), which is similar to SELV except that it has
a protective earth connection for functional reasons. (c) Functional ELV (FELV), any ELV circuit that
does not fulfill SELV or PELV requirements.
4 Ripple-free normally means a rms ripple voltage not more than 10% of the DC voltage.
80 An Introduction to Safety Grounding

TABLE 5.9
Maximum Permissible Duration of Touch Voltages
Prospective Touch Voltage
Maximum Breaking Time
of Protective Device AC (rms)* DC
(s) (V) (V)
∞ <50 <120
5 50 120
1 75 140
0.5 90 160
0.2 110 175
0.1 150 200
0.05 220 250
0.03 280 310

*root mean square

Other letters: these indicate the arrangement of the neutral and protective
conductors,
C: neutral and protective conductors combined as a common single conductor.
S: neutral and protective functions provided by separate conductors.

Types of grounding systems


By system here is meant a single source and all the installations which it supplies.
TN system: In this system the source has one (or more) ground point directly
grounded and the exposed conductive parts are connected to that point by a protec-
tive conductor. There are three variants of this system:

TN-C: In this system the neutral conductor is also the protective conductor
throughout the system.
TN-S: This system has a separate protective conductor throughout the system.
TN-C-S: In this system the neutral conductor is used as protective conductor only
in one part of the system.
TT system: The supply has a single earthed point and exposed conductive parts
are grounded through an independent grounding system.
IT system: In this system the supply is isolated from earth or connected to earth
through a high impedance and all exposed conductive parts are connected to an earth
electrode.
Figure 5.27 shows simplified circuits of the above-mentioned systems.

5.5.1 the tn-c SyStem


Figure 5.28 shows several loads fed from a distribution system whose neutral conduc-
tor is grounded at the source (transformer). If a fault occurs between a live conductor
and a metal enclosure (point 1), the neutral conductor ensures the return path for the
Grounding Systems 81

source utility consumer load


L
TN-C
N &PE

L
TN-S
N
PE

L
TN-C-S
N
PE

L
TT
N
PE

L
IT
N
PE

FIGURE 5.27 Different types of grounding systems.

Neutral
2 2'

protective
device

1
metal enclosure

FIGURE 5.28 TN-C grounding system.


82 An Introduction to Safety Grounding

fault current thus satisfying requirement A. Should a break occur in the neutral con-
ductor at point 2 or 2΄, then the potential of the enclosure will rise to that of the live
conductor when the load is switched on, thus exposing the user to grave danger. It is
evident that this arrangement does not meet the necessary safety requirements and
therefore must not be used.

5.5.2 the tt SyStem


In this system a separate grounding point must be provided locally for each consumer
(Figure 5.29). If a short circuit occurs between a live conductor and any conduct-
ing enclosure (point 1 in figure), the earth acts as a return conductor. To ensure that
the permissible touch voltage Vt is not exceeded, the total resistance Rc + Rn of the
ground fault circuit must not exceed a critical value Rt given by

Vt V
Rt ≤ ≤ t
I a kI N
where
Vt = permissible touch voltage. Since fast-acting protective devices are not always
available for small consumers, a maximum of 50 V has been set for this voltage (see
Table 5.9),
Ia = trip current (within 5s) of protective device,
IN = rated current of protective device,
K = constant whose value depends on the type of protective device used,
= 3 for fuses,
= 1.5 for overcurrent circuit breakers.5
If we assume that, Vt = 50 V, IN = 50 A, k = 3, we find that

Rt ≤ 1
3 Ω

Rn

Rc

FIGURE 5.29 TT grounding system.

5 These values for k are typical. The actual value for a particular protective device used can be found
from the tripping characteristics of that device as supplied by its manufacturer.
Grounding Systems 83

Since it is practically very difficult to achieve such low values for the resistance to
ground without excessive costs, this system of grounding is only economical when
used in conjunction with current-operated earth leakage circuit breakers (ELCB),
also known as residual current circuit breakers. The operating principle of these
breakers is as follows (Figure 5.30). The live and neural conductors supplying single-
phase loads, and all live and neutral conductors supplying three-phase loads, pass
through a toroidal magnetic core onto which is wound a secondary coil connected to
the trip circuit. Under normal operating conditions the sum of the currents linking
the core is zero and hence no magnetic flux is generated in the core. If there is an
earth fault or earth leakage on the load side of the breaker the sum of the currents
which link the core is not zero and the net current generates a magnetic flux in the
core which in turn induces a voltage in the secondary coil. If the net current is equal
to or greater than the rated current of the breaker the tripping circuit disconnects
the breaker. The value of the rated leakage current varies between 10 and 500 mA
and is chosen according to the type of load and nature of the environment. Regions
of high humidity or high pollution require higher ratings than other regions. For the
majority of domestic applications the rated current is 30 mA, and for this current, the
total resistance to ground Rc may be as high as 1,670 Ω. The operating time of earth
leakage breakers varies between 10 and 30 ms.

L N

I I –ΔI
earth leakage CB
ELCB

operating time
resistor 30 –10ms

ΔI

Rc
test circuit
L1 L2 L3 N

magnetic core

to three-phase equipment

FIGURE 5.30 ELCB.


84 An Introduction to Safety Grounding

Because of the high sensitivity of these breakers to any leakage current, they may
be used as a protection against the possibility of fire breaking out due to excessive
leakage current. The heat generated at the point of fault by the passage of currents
smaller than the rated breaker current is insufficient to ignite a fire. These breakers
require routine checking and the majority have a self-test circuit incorporated in the
design.
ELCBs protect against earth faults but do not provide protection against over-
current. There are breakers, however, which provide a combination of overcurrent
(thermomagnetic) and earth leakage protection (ELCBO).

5.5.3 the tn-S SyStem


In this system five conductors are used instead of four (Figure 5.31). The fifth con-
ductor is the protective conductor and is designated either as G (ground) or as PE.
All exposed metal enclosures and equipment parts which have to be grounded
are connected to the protective conductor. This conductor provides the return path
for the short-circuit current when an earth fault occurs between a live conductor
and the metal enclosure (point 1 in figure); the current returns to the source through
this conductor instead of through the neutral as in the TN-C system (Figure 5.28) or
through the earth as in the TT system (Figure 5.29). In this system the only danger
to a customer is if there is a break in the protective conductor (point 3 or 3΄) when an
earth fault occurs at point 1.

(a) Neutral
N 2
PE(G) Protective Earth
3
3'

R2 R1

If fuse or CB earthed
(b) Rs enclosure

PE
R gc

Vtouch

FIGURE 5.31 (a) TN-S system; (b) earth fault with touch voltage Vtouch = If Rgc.
Grounding Systems 85

Figure 5.31b shows the equivalent circuit for a fault between the live conductor
and the metal enclosure. The short circuit current is

V
If =
( X s + X 2 ) + ( R1 + R2 + Rgc )2
2

where
V = source voltage
Xs = inductive reactance of source
X2 = inductive reactance of supply conductor
R1 = resistance on conductor from service entrance to fault
R2 = resistance on conductor from source to service entrance
Rgc = resistance of protective conductor from equipment to earthing point.

The reactance of the source can be neglected so can that of cables whose cross section
is not greater than 35 mm2 (the reactance of a 35 mm2 conductor of a three-core cable
is 0.092 Ω/km while its resistance is 0.627 Ω/km) so that the fault current is

V
If =
R1 + R2 + Rgc

Now the magnitude of the touch voltage, which is the voltage to which the enclosure
will rise under fault conditions, is Vtouch = If Rgc, and its duration is determined by the
time/current characteristics of the protective device.
For miniature circuit breakers with instantaneous tripping (Figure 5.32) the
disconnection time does not exceed 0.1s (a tripping time less than 0.1s is considered
instantaneous). This time is within that specified by international regulations which

0.1s

IN Ii I

FIGURE 5.32 Time/current characteristic of miniature circuit breaker with instantaneous


tripping.
86 An Introduction to Safety Grounding

require that for circuits supplying fixed equipment the maximum disconnection time
is 5s, whereas for circuits supplying socket outlets the disconnection time shall not
exceed 0.4 s. The reason for this is that socket outlets may be used to supply a hand-
held equipment requiring that it be gripped continuously in operation (e.g., hand
drill) which constitutes a much greater risk to the user than fixed equipment. Now
setting

I f = a I N = Ii

where
IN = rated breaker current
Ii = instantaneous tripping current
and for Vtouch = 50 V, we have that

50
Rgc ≤
Ii
Having determined Rgc and the fault current If , the time t for disconnection of this
current is determined from the relevant time/current characteristics of the breaker
used. Substitution for If , t, and the appropriate value of k in Eq.(5.2) gives the mini-
mum cross-sectional area of the protective conductor and this has to be equal to or
less than the size determined by its resistance Rgc.
Figure 5.33 shows the neutral (N) and protective (PE) conductors for single-phase
and three-phase supply systems when the premises are fed from a general distribu-
tion system and when fed from a transformer on the premises. In all cases one must

service entrance
service entrance

N
N
PE
PE
only connection
permitted between N
and PE

transformer on premises

PE

FIGURE 5.33  Connection between neutral and protective conductors.


Grounding Systems 87

make sure that there are no other connections between the neutral and protective
conductors on the load side of the supply. The reason for this is that should there be
another connection, then one part of the load current will return through the protec-
tive conductor. This is absolutely forbidden during normal operation because in this
case the potential of the metallic parts connected to the protective conductor will rise
to a value equal to the voltage drop in that conductor (current × resistance).
It is possible to use the metallic covering (lead sheath and armour) of the cable
supplying the installation as the protective conductor. In this case it must be ensured
that all junctions are capable of carrying the current flowing in the cable covering. If
nonmetallic junction boxes are used the electrical continuity of the covering must be
ensured by means of jumpers whose resistance must not exceed that of the covering
which has been removed. If the cable is only armored, then in the majority of cases
(except if the armor has copper wires) it is difficult to obtain an armor resistance
which is sufficiently low to allow the return current to operate the consumer’s
protective device when an earth fault occurs.

5.5.4 protective multiple earthing


This is a modification of the TN-C system. In this system also the neutral conductor
is used as the return conductor for the fault current (Figure 5.34). However, to avoid
the drawbacks of the TN-C system the neutral conductor is earthed at the trans-
former, at the end of the feeder, and at several points along its length (about three
points per kilometer) such that the resistance of the neutral conductor between any
point and ground does not exceed 10 Ω (or any other value specified by the Authority
responsible for the distribution).
It is evident that a break in the system neutral conductor (point 2 in figure) does
not constitute a danger to the consumer in the event of a short circuit between a
live conductor and a metal enclosure (point 1) since in this case the ground acts
as a return conductor for the fault current. However, if a break occurs in the con-
sumer neutral conductor (point 2΄) then under fault conditions the metal enclosure’s
potential will rise to that of the live conductor when the equipment is switched on.
To avoid this it is preferable to use a separate protective conductor connected to a

2'

FIGURE 5.34 PME system.


88 An Introduction to Safety Grounding

consumer installation earth which is electrically independent of the source earth


(Figure 5.35). In this case there is no danger to the consumer except when a double
fault occurs: a break at 2΄ and a short circuit at 1. However, a break at 2΄ in the PE
conductor is much less likely to occur than a break at 2΄in the neutral connection in
Figure 5.34.
It should be stressed that under no circumstances should a direct earthing system
be used with any PME system. The following example shows why.
When a fault occurs at point 1 in Figure 5.36, a current of 230/11 = 21 A flows
to ground and the potential of the neutral conductor rises to 21 × 10 = 210 V. This
potential appears on all metal bodies connected to the neutral conductor.

2
N

N
PE
2'

FIGURE 5.35 Use of separate PE conductor in the PME system.

400/230 V

R = 10 Ω
PME equivalent
earth resistance
1

20 V 200 V
R=1Ω

FIGURE 5.36 Danger of combining TT and PME systems.


Grounding Systems 89

5.5.5 the it SyStem


In this system the neutral point at the source is either completely isolated or grounded
through a high resistance (impedance) and exposed equipment parts are grounded
locally as shown in Figure 5.37. The IT system is only used where continuity of
supply is of vital importance such as in some chemical plants with continuous
processes and in some medical applications as well as when an extremely low value
of first earth fault current is required. In this system the equipment is grounded
either through a common protective conductor (Figure 5.37a) or through an indi-
vidual grounding arrangement (Figure 5.37b). In both cases a first fault between a
conductor and ground (earthed enclosure) does not affect the continuity of supply.
Should a second fault occur between another conductor and ground the protective
device must disconnect the faulted circuit. It is preferable that the protective device
is an ELCB backed by fuses and it must be made certain that the touch voltage does
not exceed 50 V.
For such systems the majority of international specifications recommend the
installation of an earth leakage monitor. This device monitors the condition of the
insulation resistance between the isolated system and ground; if this resistance falls

(a) Fault current when a second earth fault


occurs in addition to the first fault

earth leakage
monitor
7 7 second fault

(b)

earthleakage
monitor
7 7

local earthing

FIGURE 5.37 The IT system: (a) using a common PE conductor; (b) using individual
grounding conductor.
90 An Introduction to Safety Grounding

below a preset value (between 50 and 200 k·Ω depending on the nature of the instal-
lation), the device sets off a sound alarm to warn those responsible of the existence
of a fault so that it may be located and removed without having to disconnect the
supply.

5.6 PROTECTION BY ISOLATION


The load is isolated from the supply source by an isolation transformer. This is a
transformer with two completely isolated windings and a 1:1 transformation ratio.
These transformers are used for supplying power to portable equipment, especially
hand-held tools, to protect users should an earth fault occur between a live conductor
and the casing. Figure 5.38a gives an example of the use of an isolation transformer.
If the equipment is being used on a metal platform or scaffolding then, in order to
prevent the passage of current through the body of the operator in case of a double
fault, the equipment casing must be connected to the metal structure as shown in
Figure 5.38b.
When using equipment in a confined metal enclosure it is not permitted to supply
more than one load from the same transformer and the rated load must not exceed
16 A. When more than one load is supplied by an isolation transformer and there is
the possibility of touching both frames at the same time, then the frames have to be
connected together as shown in Figure 5.39.

Double insulation is the use of additional insulation over and above the basic insula-
tion required. A common example of double insulation is wire with an insulating

1:1 isolation
transformer
(a)
7

(b)
safety connection

metal structure

FIGURE 5.38 Use of an isolation transformer as a safety measure.


Grounding Systems 91

isolation
transformer

tie connection between casings


(not grounded)

FIGURE 5.39 Isolation transformer supplying more than one load.

cover placed inside a second insulating cover. For electrical machinery and equip-
ment double insulation is accomplished either during the manufacturing or assem-
bling stages or by placing it before use inside an insulated enclosure. Equipment
provided with double insulation must conform with the norms and standards appli-
cable to such equipment.

5.7 GROUNDING OF COMPUTERS AND


DATA PROCESSING EQUIPMENT
For safety purposes all standards recommend the grounding of all exposed metal
casings and enclosures as well as the signal reference plane (SRP) and that there
be only a single common ground point. Figure 5.40 gives the grounding elements
of a computer or data processing center although it must be emphasized here that
the grounding of electronic equipment in particular is coupled with electromagnetic
interference (compatibility) and as such is a subject on its own. For additional
information on this topic the reader may consult references [17] and [31].

equipment

grounding connection
provision
equipment
e

to common
ground signal reference plane

equipment

FIGURE 5.40 Grounding elements for a computer center.


6 Substation Grounding
Systems

6.1 INTRODUCTION
Substations are an essential part of any electrical power grid and their grounding
system must be designed to ensure a safe environment for people who happen to be
in and around the substation during a ground fault. This is achieved by limiting the
ground potential rise (GPR)—and hence the resulting step and touch voltages—to
a value that is below the hazard level for human beings. For large substations the
resistance of the grounding system should not exceed 1Ω while for smaller stations
it should not exceed 5 Ω.
The basic substation ground system used by most utilities1 is a grid system
consisting of bare copper cables buried horizontally about 30–40 cm in the ground
and forming a network of squares or meshes (Figure 6.1). Such a system usually
extends over the entire substation area. The spacing of the conductors (mesh size)
varies according to the requirements of the installation.
A grid is used instead of the conventional driven rods for the following reasons:

• rods
– wire

FIGURE 6.1 A typical meshed ground grid.

1 The outline given in this chapter is based essentially on IEEE Std.80-2000: Guide for safety in AC
substation grounding.
93
94 An Introduction to Safety Grounding

form a grid or mesh network which in itself constitutes a grounding system.


In a soil of reasonably low resistivity it has been found that this network may
be so effective as to make the original driven rod electrodes unnecessary.

Connections between the various ground leads and the wire grid as well as
connections at the crossovers within the grid are usually clamped, brazed, or
welded. Crossovers should be brazed or welded. Ordinary soldered connections
are to be avoided because of failure under high fault current or because of galvanic
corrosion. Each element of the ground system has to be designed so as to resist
fusing and deterioration of electric joints under the worst ground fault conditions
using Eq.(5.1).
A typical grid system usually extends over the entire substation area and sometimes
beyond the fence which surrounds the building and equipment. Since the fence is
usually located on the periphery of the grid area where surface potentials are highest,
the grounding of the fence is of major importance especially since the outside of the
fence is usually accessible to the general public. It is therefore preferable to extend
the area of the ground mat such that the fence lies within it. Readers should consult
Ref. [7] for a full discussion of fence grounding.

6.2 DESIGN
 STEPS
The design engineer will draw up a preliminary design for the grid and then check
that it meets the safety requirements. If not, the design shall be modified (repeat-
edly if necessary) until all safety requirements are satisfied. The design involves the
following steps:

1. Determination of soil resistivity


2. Computation of mat resistance to ground
3. Computation of maximum potential rise
4. Computation of touch and step voltages
5. Ensure that safety requirements are satisfied

6.2.1 ground mat reSiStance


The ground mat of a substation may be approximated by a circular disc whose area is
equal to that of the substation (Figure 6.2). The resistance of the disc can be obtained
from Eq.(3.5) given in Section 3.4, for h = 0,

ρ
R= (6.1)
4a

where
R = resistance of mat to ground (Ω)
ρ = average ground resistivity (Ω·m)
a = radius of disc of same area as that occupied by mat (m)
Assuming that the mat area is A m2, the radius of the disc is
Substation Grounding Systems 95

FIGURE 6.2 Representation of a ground mat by an equivalent disc.

A
a=
π
so that
ρ π
R= (6.2)
4 A
This equation is usually modified by adding a second term as follows:

ρ π ρ
R= + (6.3)
4 A L

where L is the total length of buried conductors in meters. This second term takes
into consideration the fact that the ground mat resistance is greater than that of a
solid circular disc and that this difference decreases as the length of the conductor
increases (for a solid disc L = ∞).
Equations (6.2) and (6.3) are approximate and there are more accurate alternative
formulas for estimating the ground mat resistance. Sverak’s formula2 which takes
into account grid depth is

1 1  1 
R = ρ + 1+  (6.4)
L 20 A  1 + h 20 / A  

where h is the depth of the ground mat, usually between 0.25 and 2.5 m.

Example 6.1

Figure 6.3 shows a ground mat whose area is 60 × 60 m.


The ground mat consists of 4 × 4 m meshes and is buried at a depth of 0.3 m.
The conductor size is 2/0AWG (133100 cmil) copper. The soil has a uniform
resistivity of 100 Ω·m. It is required


(b) to determine the effect of mat area and burial depth on the mat resistance
to ground.

2 J.G.Sverak,Simplified analysis of electrical gradients above a ground grid; Part I—How good is the
present IEEE method? IEEE Trans.Power Apparatus Syst. vol. PAS-103, no. 1, pp.7–25, 1984.
96 An Introduction to Safety Grounding

(a) We have that √A = 60 m, ρ = 100 Ω·m, h = 0.3 m


Total number of conductors = 32
Total length of wire L = 32 × 60 = 1,920 m.
From Eq.(6.2) the mat resistance to ground R = 0.7384
From Eq.(6.3) the mat resistance to ground R = 0.7890
From Eq.(6.4) the mat resistance to ground R = 0.7970

The mat resistance to ground for different values of mat area and burial depth has
been calculated from Eq.(6.4) and the results are plotted in Figures 6.4 and 6.5.
The conclusions are as follows:
• The resistance varies almost inversely as the square root of the mat
area.
• The resistance decreases slightly with increasing burial depth.

4m
4m

FIGURE 6.3 Ground mat divided into meshes 4m × 4 m.

2.5

2.0
resistance, ohm

1.5

1.0

0.5

0
0 1 2 3 4 5 × 103
area, m2

FIGURE 6.4 Effect of area on mat resistance.


Substation Grounding Systems 97

0.9

resistance, ohm
0.8

0.7

0.6

0.5
0 0.5 1.0 1.5 2.0 2.5
burial depth , m

FIGURE 6.5 Effect of burial depth on mat resistance.

6.2.2 calculation of maximum gpr


To ensure the safety of personnel the ground mat must be designed so as to limit the
GPR of the substation ground to an acceptable value when a short circuit to ground
occurs. Determination of this GPR is quite complicated because it depends on many
factors which include type of substation, location of fault, magnitude of fault current,
and grounding system resistance. The maximum GPR can be calculated as follows:
(GPR)max = I G Rg (6.5)

where
Rg = resistance of mat to ground.
IG = it is the greatest value of rms current which flows from the ground grid to
the surrounding ground. It is this current which produces the greatest rise in ground
potential. The value of IG is defined as follows:

IG = C f D f S f I f (6.6)

where
If = rms value of the symmetrical ground fault current;
Cf = correction factor for future system growth (>1);
Df = decrement factor which takes into consideration the attenuation due to the
passage of fault current from an asymmetrical to a symmetrical current:

Time, s Df
0.01 1.65
0.1 1.25
0.25 1.10
≥ 0.5 1.00
98 An Introduction to Safety Grounding

For intermediate values of fault duration, decrement factors may be obtained by


linear interpolation.
Sf = current division factor. It represents the ratio between the current which
actually flows from the grid to ground, and the short-circuit current and its value are
usually between 50% and 80% depending on the location of the fault.
Figure 6.6 gives examples of the influence of fault location on the flow path of the
fault current.
In the examples shown in Figure 6.6 we note the following:

In (a) the station neutral is the only point connected to ground. Almost the
whole of the fault current flows in the ground mat.
In (b) the neutral point is connected to ground outside the station and the entire
fault current flows from the mat to ground.
In (c) there is another grounded neutral point outside the station. One part of
the fault current flows to ground.
In (d) the fault current divides between the grounded points according to the
ground path resistances.

6.2.3 determination of maximum touch and Step voltageS


The two most important parameters in the design of a substation ground mat are the
touch and step voltages. Touch voltage in a grid ground is defined as the difference in
potential between the grounded structure and the center of a grid mesh. The highest
value of this voltage appears at the center of a mesh at a perimeter corner and is
referred to as the “mesh voltage.”
The maximum touch (mesh) and step voltages can be determined using the
following formulas,3


Emesh = ρ K m K i I G / L (6.7)

Estep = ρ K s K i I G / L (6.8)

Emesh = ρ K m K i I G / ( L + Lr ) (6.9)

Estep = ρ K s K i I G / ( L + Lr ) (6.10)

Emesh = ρ K m K i I G / ( L + 1.15Lr ) (6.11)

Estep = ρ K s K i I G / ( L + 1.15 Lr ) (6.12)

3 For a mathematical analysis of the gradient problem in a ground grid the reader should consult
IEEE Std 80-2000: Guide for safety in AC substation grounding.
Substation Grounding Systems 99

(a) Fault current follows metallic path (b) Total fault current
provided by ground grid. flows from station ground
No appreciable current flow in earth. grid to earth.

other system
grounds

(c) Fault current returns to local neutral through local


grid and to remote neutrals through earth. The latter
current is of concern in study of danger voltages.

local station ground other system grounds

(d) Portion of fault current returning from earth to ground grid at


local station determines grid potential rise and gradients there.

FIGURE 6.6 Fault location and path of current flow: (a), (b), and (c) fault within local station;
(d) fault outside station. (Reproduced with permission from IEEE Std. 80 Guide for Safety in
AC Substation Grounding.)


100 An Introduction to Safety Grounding

where
ρ = resistivity of soil in ohm-m.
L = total length of all grid conductors (m).
Lr = total length of the ground rods (m).
Ki = nonuniformity factor which accounts for the nonuniform ground current flow
from different parts of the grid.
Ks, Km = coefficient which take into account the effect of number, spacing, size,
and depth of burial of the grid conductors.

The values of Ki, Ks, and Km are determined from the following formulas:
K i = 0.656 + 0.172n (6.13)
1   D2 ( D + 2h) 2 h  K ii 8 
Km =  ln  + −  + ln  (6.14)
2π   16hd 8 Dd 4 d  K h π(2n − 1) 

1 1 1 1
Ks = + + (1 − 0.5n − 2 )  (0.25m < h < 2.5m) (6.15)
π  2h D + h D 

For grids with no ground rods or with evenly spaced rods,

1
K ii = (6.16)
(2n)2 / n

For rods along the perimeter,

K ii = 1

For all cases,

K h = 1 + h / h0 (6.17)

n = number of parallel grid conductors in one direction,


D = spacing between parallel conductors (m),
h = burial depth (m),
ho = 1 m (reference depth).

Figure 6.7 shows the voltage profile on the surface of the ground along a diagonal line
above a ground mat with a 10 × 10 mesh and mesh size 6 m × 6 m. If a person stand-
ing at point P, which is at the center of the corner mesh, is in touch with a grounded
part during the occurrence of a ground fault, he will be subjected to the touch voltage
Vtouch; a person standing at a corner with his feet 1 meter apart (F1F2) will be sub-
jected to a step voltage Vstep as shown.
Figure 6.8 shows the effect which the number of meshes into which a grid is
divided has on the touch voltage. It can be seen that the maximum touch voltage
(mesh voltage) decreases as the number of meshes increases.
Substation Grounding Systems 101

F2
F1
P

(0,0) centre of mesh

GPR

100 Vtouch
voltage (% of GPR)

Vstep
80

60

0 6 12 18 24 30 X,Y

FIGURE 6.7 Voltage profile along the diagonal of a ground mat (10 × 10 mesh, size 6 m ×
6 m, 2/0 AWG copper conductors, burial depth 0.3 m).

Y Y Y

X X X

(a) (b) (c)

100
(c)
voltage (% of GPR)

(b)
80

(a)
60

0 15 30 45

FIGURE 6.8 Effect of mesh size on touch voltage.


102 An Introduction to Safety Grounding

6.2.4 Safety requirementS


After calculating the values of the maximum touch (mesh) and step voltages from
Eqs. (6.7) to (6.12) we must check that these values do not exceed those permitted
(see Chapter 4) viz.,

(116 + 0.17ρ )
Vtouch (max.permitted) = (6.18)
t

(116 + 0.7ρ )
Vstep (max.permitted) = (6.19)
t

or if the ground surface has been covered with a layer of high resistivity such as
crushed rock,

116 + 0.17Cs ρs
Vtouch (max) = (6.18a)
t

116 + 0.7Cs ρs
Vstep (max) = (6.19a)
t

where Cs is a resistivity derating factor as described in Section 4.2. Thus the safety of
the grounding system must satisfy the following conditions:

Vtouch (max) ≥ Emesh (6.20)

Vstep (max) ≥ Estep (6.21)

Example 6.2

For the ground mat of Example 6.1 if the symmetrical ground fault current is 1,000
A and the duration of the fault is 0.5 s, it is required to determine the following
voltages:
(GPR)max , Emesh , Estep , Vtouch (max), Vstep (max)

Depth of burial h = 0.3 m, ρ = 100 Ω·m, diameter of 2/0 AWG conductors


d = 0.0106 m. Assume Sf = 0.8 and Cf = 1.5.
From Figure 6.3 we see that
n = 16, D = 4 m, L = 1920 m
From Example 1 the resistance to ground of the mat was found to be Rg = 0.79
Ω so from Eq.(6.5) we have that
(GPR)max = 1200 × 0.79 = 947 V
and from Eqs.(6.7) and (6.8) we find that
Emesh = 152.24 V
Estep = 145.7 V
Substation Grounding Systems 103

and from Eqs. (4.4) and (4.3) we find that the maximum permissible touch and
step voltages are
Vt (max) = 189 V
Vs (max) = 262 V
Since the two conditions (6.20) and (6.21) are met, no modifications to the
mesh are necessary.

Example 6.3

Assume that the mesh size of the mat given in Example 6.2 is 10 × 10 m and all
other details remain the same. It is required to find the touch and step voltages
and introduce any necessary modifications required to make these voltages safe.
From the appropriate equations we find that

Emesh = 295V, Estep = 158 V

Vt ( max ) = 189 V,Vs ( max ) = 262 V

The touch voltage does not satisfy the safety requirement. It is possible to introduce
two modifications in the design:

rs

Emesh = 157 V < Vt ( max ) = 189 V

Estep = 84 V < Vs ( max ) = 262V

Vt ( max ) = 465V > Emesh = 295V

Vs ( max ) = 1,402V > Estep = 158V

The correction factor Cs used in Eqs.(6.18a) and (6.19a) is taken from Figure 4.3 for
hs = 0.1m and K = – 0.9.
104 An Introduction to Safety Grounding

6.3 DESIGN OF GROUND MAT USING COMPUTER PROGRAM


The reader will realize from the foregoing that the design of ground mats requires
laborious calculations. However, there are a number of computer programs specially
designed for this purpose and we give here an example of mat design using one such
program (EDSA) which uses the Sverak formula as given by Eq. (6.4).
Figure 6.9 shows the preliminary design of a substation ground mat of area 150 ×
150 ft. The size of the copper conductors is 2/0 AWG and the mat is buried at a
depth of 1.5 ft in soil of uniform resistivity 100 ohm-m. The fault current is 1,000
A and its duration 0.25 s. It is required to determine from the program if the safety
requirements are met in the following cases:

Tables 6.1–6.3 give the respective computer outputs for these three cases:

(A) Table 6.1 shows that safety requirement are met since Emesh and Estep are less
than Vt (max) and Vs (max).
(B) Table 6.2 shows that 20 driven rods are required to meet the safety requirements.
(C) Table 6.3 shows the large increase in Vt (max) and Vs (max) brought about by
the using a layer of crushed rock.

15 ft
15 ft

150 ft

150 ft

FIGURE 6.9 Ground mat of area 150 ft2 with 10 × 10 meshes, burial depth 1.5 ft.
Substation Grounding Systems 105

TABLE 6.1
Job: Solution of Example (Case A) Ground Grid Design using
EDSA Program EE Dept. College of Engineering
Grid Data
Grid length 150.00ft Grid width 150.00ft
Conductor spacing 15.00 ft Ambient temperature 77.00 F
Sym. grnd. fault 1,000 amps Fault duration 0.25 s

Current Division Factor 0.80

Resistivity
Soil 100.00 Ω·m Crushed rock 2,500.00 Ω·m
Concrete 0.00 Ω·m

Depth of Burial
Grounding grid 1.50 ft Crushed rock 0.00 ft
Reference depth of grid 3.28 ft Concrete radius 0.00 ft

Ground Rod Information


Number Length (ft) Diameter (in.)
20 24.61 0.748
36 32.81 0.748
36 49.21 0.748
38 32.81 0.748
48 39.37 0.748

Thermal
Material Conductivity Factor Oc FusingTemp. Resistivity Capacity
STDANNEALED-SOFT-CU 100.00 0.003930 234.0 1,083.0 1.724 3.422
106 An Introduction to Safety Grounding

AWG/MCM Cir.Mil Cond.Dia CM Con.Dia IN


2/0 AWG 133,100 1.0600 0.4190

Output Results
Voltages
V (max. touch) 268.33 V (max. step) 377.32
E (mesh) 171.06 E (step) 107.19

Grid potential rise 929.84

Grid Particulars
Conductor length 3,300.00 ft
Max grid current 880.00 amps
Grid resist 1.057 Ω

Material AWG/MCM Cirmil Grid Cond Dia (in.)


STD-ANNEALED-SOFT-CU 2/0 AWG 133,100 0.4173

*** No rods are necessary ***

TABLE 6.2
Job: Solution of Example (Case B) Ground Grid Design using EDSA
Program EE Dept. College of Engineering
Grid Data:
Grid length 150.00 ft Grid width 150.00 ft
Conductor spacing 30.00 ft Ambient temperature 77.00 F
Sym. grnd. fault 1,000 amps Fault duration 0.25 s

Current division factor 0.80


Substation Grounding Systems 107

Resistivity:
Soil 100.00 Ω·m Crushed rock 2,500.00 Ω·m
Concrete 0.00 Ω·m

Depth of burial:
Grounding grid 1.50 ft Crushed Rock 0.00 ft
Reference depth of grid 3.28 ft Concrete Radius 0.00 ft

Ground Rod Information


Number Length (ft) Diameter (in.)
20 24.61 0.748
36 32.81 0.748
36 49.21 0.748
38 32.81 0,748
48 39.37 0.748

Fusing Thermal
Material Conductivity Factor Oc Temp Resistivity Capacity
STD-ANNEALED-SOFT-CU 100.00 0.003930 234.0 1,083.0 1.724 3.422

AWG/MCM Cir.Mil Cond.Dia CM Con.Dia IN


2/0 AWG 133,100 1.0600 0.4190

Output Results
Voltages
V (max. touch) 268.33 V(max. step) 377.32
E (mesh) 192 E (step) 85.25
Grid potential rise 1,095.16
108 An Introduction to Safety Grounding

Grid Particulars
Conductor length 1,800.00 ft
Max grid current 880.00 amps
Grid resist 1.244 Ω

AWG/ Grid Cond Ground Length of Dia.of


Material MCM Cirmil Dia (in) Rods Each Rod Rod
STD-ANNEALED-SOFT-CU 2/0 AWG 133,100 0.4173 20 24.61 ft 0.748 in.

TABLE 6.3
Job: Solution of Example (Case C) Ground Grid Design using
EDSA Program EE Dept. College of Engineering
GRID DATA:
Grid length 150.00 ft Grid width 150.00 ft
Conductor spacing 15.00 ft Ambient temperature 77.00 F
Sym. grnd. fault 1,000 amps Fault duration 0.25 s

Current division factor 0.80

Resistivity:
Soil 100.00 Ω·m Crushed rock 2,500.00 Ω·m
Concrete 0.00 Ω·m

Depth of Burial:
Grounding grid 1.50 ft Crushed rock 0.33 ft
Reference depth of grid 3.28 ft Concrete radius 0.00 ft
Substation Grounding Systems 109

Ground Rod Information


Number Length (ft) Diameter (in)
20 24.61 0.748
36 32.81 0.748
36 49.21 0.748
38 32 81 0.748
48 39.37 0.748

Fusing Thermal
Material Conductivity Factor Oc Temp Resistivity Capacity
STD-ANNEALED-SOFT-CU 100.00 0.003930 234.0 1,083.0 1.724 3.422

AWG/MCM Cir.Mil Cond.Dia. CM Con.Dia. IN


2/0 AWG 133,100 1.0600 0.4190

Output Results
Voltages
V (max. touch) 712.58 V (max. step) 2,154.31
E (mesh) 155.51 E (step) 97.44
Grid potential rise 845.31

Grid Particulars
Conductor length 3,300.00 ft
Max grid current 800.00 amps
Grid resist 1.057 ohms

Material AWG/MCM Cirmil Cond Dia (in)


STD-ANNEALED-SOFT-CU 2/0 AWG 133,100 0.4173

*** No rods are necessary ***


7 Static Electrification

7.1 INTRODUCTION
Static electrification is a term applied to all mechanical operations which result in
the separation of positive and negative charges. These operations include friction,
contact, or impact between two solid surfaces, a solid surface and a liquid or gas as
well as the separation of surfaces and the rupture of liquids by spraying or bubbling.
Electrification by friction or contact is generally known as tribo-electrification or
contact electrification and the static electrification produced by spraying (atomizing)
or bubbling of liquids is known as spray electrification. The electrification of dusts
and powders is a type of solid-solid tribo- or contact-electrification. It should be
mentioned here that experimental results have confirmed that tribo-electrification
is an effect due to contact between two surfaces and their subsequent separation;
friction has no effect on the electrification process unless it leads to an increase
in the temperature of one surface above that of the other. The phenomena of static
electrification are of major importance in industry and can cause explosions in sugar
factories, granaries, sulfur mills, explosives factories, petrochemical factories, and in
all branches of the petroleum industry as well as in operations involving the handling
of coal and inflammable liquids. Moreover, static electrification has been known to
cause explosions in the operating rooms in hospitals.
With the great advances in the manufacture of integrated circuits and their use
in all modern electronic devices and systems, electrostatic discharge (ESD) con-
stitutes a serious source of damage to these circuits during their manufacture or
their packaging, as well as during their operation. Experience has shown that such
discharges lead to the destruction of thin film metal oxide semiconductors and a
number of sensitive elements such as tracks, membrane resistances, and capaci-
tors in integrated circuits, etc. They may also cause the so-called ESD latent static
damage in which an initial discharge will cause only partial damage to a track
leading to its gradual degradation and eventual failure due to thermal or vibrational
stresses during use.
Before considering the subject of static electrification and ways to control it, we
have found it pertinent to acquaint the reader with a synopsis of the conditions which
are necessary for an ESD to cause a fire or an explosion.

7.2 CONDITIONS NECESSARY FOR IGNITION


Burning of a fuel in air is the result of the reaction between the fuel molecules and the
oxygen molecules in the air. For liquid fuels the mixture of fuel and oxygen becomes
available when the fuel evaporates, but for solid fuels there must be a break in the
chemical bond of the fuel molecules so that the resulting hydrocarbon molecules can
interact with the oxygen.

111
112 An Introduction to Safety Grounding

A fire or explosion will occur when the following three elements exist together:
fuel, oxygen, and an ignition source; these constitute the so-called danger triangle.
However, the simultaneous presence of these elements per se will not lead to a fire or
explosion unless the following conditions prevail:

7.2.1 exploSion limitS


For any kind of fuel, whether gaseous, liquid, or solid, self-sustained combustion
is only possible for volume concentrations between a lower and upper limit known
as the explosion or flammability limits (EL or FL) which must be determined by
experiment.
The lower explosive limit (LEL) is the lowest concentration of a gas or vapor in
air which will burn or explode in the presence of an ignition source. Below the LEL
the mixture is “too lean” to burn (i.e., there is insufficient fuel). The upper explosive
limit (UEL) is the highest concentration of a gas or vapor in air which will burn or
explode in the presence of an ignition source. Above the UEL the mixture is “too
rich” to burn (i.e., there is insufficient oxygen).
For gases and vapors both these concentrations are expressed as the percentage
(volume) concentration of fuel in the fuel/air mixture (in oxygen-rich atmospheres,
such as hospitals, concentrations should be known for fuel/oxygen mixtures). For
dusts (or powders) they are expressed as the mass of dust per unit volume of dust/air
mixture, e.g., kg/m3.


The ignition of the liquid fuel depends on its vapor pressure. This
pressure, and hence the vapor concentration, depends on the tempera-
ture. From the relationship between vapor pressure and temperature for
a given liquid we obtain the relationship between vapor concentration
and temperature. The temperature corresponding to the LEL is known as
the flash point of the liquid, which is defined as the lowest temperature
at which the vapor concentration in the air near the surface of the liquid
is high enough to form an ignitable mixture. The flash point of a liquid
is a measure of its flammability, the lower its flash point, and the higher
its flammability. At temperatures below their flash point liquids cannot
be ignited.
Tables 7.1 and 7.2 give the explosive limits, the flash point, and the MIE
(Sec.7.2.4) for a number of substances. The values have been chosen from
a large number of sources [38–40] and may be finely adjusted from time to
time. However, they serve as guidance and as a comparison between the
FLs and MIE of some common gases and vapors.
Static Electrification 113

TABLE 7.1
LEL and UEL and Flash Point of Some Gases and Vapors
LEL UEL Flash Point
Substance (% by vol of air) (% by vol of air) (°C)
Acetone 2.6–3 12.8–13 –17
Acetylene 2.5 82 –18
Benzene 1.2 7.8 –11
Butane 1.4 8.4 –60
Carbon monoxide 12 75 –191
Diesel fuel 0.6 7.5 > 62
Diethyl ether 1.9–2 36–48 –45
Ethanol (ethyl alcohol) 3–3.3 19 12.8
Ethylene glycol 3 22 111
Gasoline (100 octane) 1.4 7.6 < –40
Hexane 1.1 7.5 –22
Hydrogen 4 75 –
Isopropyl alcohol 2 12 12
Methane 4.4–5 15–17 –
Methanol (methyl alcohol) 6–6.7 36 11
Octane 1 7 13
Pentane 1.5 7.8 – 40 to –49
Propane 2.1 9.5–10.1 –
Toluene 1.25 6.75–7.1 4.4
Xylene 0.9–1.0 0.9–1.0 27–32

TABLE 7.2
MIE for Some Gases and Vapors
MIE Concentration in Air
Substance (mJ) (% by volume)
Acetone 1.15 4.5
Acetylene 0.017 8.5
Ammonia 680 –
Benzene 0.22 4.7
Butane 0.25 4.7
Diethyl ether 0.19 5.1
Hexane 0.24 3.8
Hydrogen 0.017 28
Methane 0.28 8.5
Methanol 0.14 –
Propane 0.25 5.2
Toluene 0.24 4.1
114 An Introduction to Safety Grounding

(ii) Dust layers and dust clouds


Dusts and powders also have lower and upper ELs but in general it is
the lower limits which are of importance. For many organic dusts these are
in the range 10–50 g/m3, which is much higher than the limit set for health
reasons. ELs also depend on the size of the dust particles involved and are
not an intrinsic property of the material. A concentration above the LEL
can be created from an accumulation of settled dust by a sudden gust of
wind. Dust accumulations can be prevented by using appropriate enclosures,
proper ventilation, and regular surface cleaning. However, the formation of
dust clouds is inherent in certain industries such as sugar mills, sulfur mills,
flour, and grain industries as well as in granaries and in the handling of coal.
Such dust clouds are explosive and some can be ignited by static sparks.

7.2.2 oxygen
Oxygen is a vital element in the combustion of any type of fuel and the majority of
fuels require a minimum oxygen concentration of about 10% by volume for combus-
tion to take place. Since under normal atmospheric conditions the oxygen content of
air is 21% by volume, this means that for absolute safety about 50% of the air must be
substituted by some inert gas such as nitrogen or carbon dioxide. This may be imple-
mented only in certain limited zones where it is absolutely necessary to guarantee a
100% explosion proof environment.

7.2.3 ignition SourceS


An ignition source is a source that is capable of providing the energy necessary to
ignite a combustible mixture. The possible sources of ignition are the following (not
listed in order of importance or of frequency of occurrence):

• Flames (cutting and welding, cigarettes, matches, etc.)


• Hot surfaces (heating pipes, electrical casings, etc.)
• Friction heating or sparks (abrasive cutting)
• Mechanical machinery
• Electrical equipment and installations (electrical sparks at make/break
switches, loose contacts, bad solder joints in cables, etc.)
• ESD sparks
• Lightning strikes
• Electromagnetic radiation of different wavelengths
• Ultrasonics
• Exothermal chemical reactions

The potential of these sources to ignite a combustible mixture varies enormously. For
example, whereas lightning and an open flame have sufficient energy to ignite any
combustible material, other sources such as mechanical or electrical sparks can only
do so if their discharge energy is equal to or greater than the MIE of the mixture.
Static Electrification 115

ignition energy
lean
MIE rich

flammable range

LEL UEL
concentration (%)

FIGURE 7.1 Variation of ignition energy with concentration of combustible material.

7.2.4 minimum ignition energy


As mentioned above, for ignition to occur, the concentration of the combustible material
must lie between the lower and upper explosion limits. For each concentration within
these limits a certain amount of energy (usually measured in milli-joules) is required to
initiate the ignition process. As we move from LEL to UEL concentrations this energy
passes through a minimum value as shown in Figure 7.1. This value represents the mini-
mum ignition energy (MIE of the mixture. Table 7.2 gives the MIE values for some
vapors and gases whereas Table 7.3 gives MIE values for dust clouds and dust layers.

7.3 GENERATION OF ELECTROSTATIC CHARGES


Surface charges are generated when two dissimilar surfaces are brought together and
then separated. During the initial contact a transfer of charge takes place between the
surfaces so that when they are separated one surface will carry a negative charge and the
other a positive charge. This method of static charge generation is known as tribo-electri-
fication although, as mentioned in the introduction, it is in effect contact electrification.
The amount of charge that is transferred from one body to another depends on
the ability of the material that donates charge or its ability to accept charge. Based
on this ability materials have been classified in a series known as the tribo-electric
series which is given in Table 7.4. Materials whose position in the table is higher than
others have a greater ability to donate electrons (and hence acquire a positive charge)
than those lying below it. Similarly materials whose position is lower than others
have a greater ability to accept electrons (and hence acquire a negative charge) than
those above them in the table. Thus if two bodies of dissimilar material come into
contact the body which acquires a positive charge is that whose position in the table
is above that of the other. For instance if a glass body comes in contact with a rubber
body, the glass will acquire a positive charge and the rubber a negative charge.
The magnitude of the acquired charge depends on a number of factors such as
the cleanliness of the surfaces, contact pressure, extent of contact area as well as the
speed with which the surfaces are separated. In many cases surface charges may
116 An Introduction to Safety Grounding

TABLE 7.3
MIE for Some Dust Clouds and Layers
Dust Cloud Dust Layer
Material (mJ) (mJ)
Alfalfa 320
Allyl alcohol resin 20 80.0
Aluminum 10 1.6
Aluminum stearate 10 40.0
Aryl sulfonyl hydrazine 20 160.0
Aspirin 25 160.0
Boron 60
Cellucotton 60
Cellulose acetate 10
Cinnamon 40
Coal, bituminous 60 560.0
Cocoa 100
Cork 35
Cornstarch 30
Dimethyl terephthalate 20
Dinitro-o-toluamide 15 24.0
Ferromanganese 80 8.0
Gilsonite 25 4.0
Grain 30
Hexamethylenetetramine 10
Iron 20 7.0
Magnesium 20 0.24
Manganese 80 3.2
Methyl methacrylate 15
Nut shell 50
Paraformaldehyde 20
Pentaerythritol 10
Phenolic resin 10 40.0
Phthalic anhydride 15
Pitch 20 6.0
Polyethylene 30
Polystyrene 15
Rice 40
Seed (clover) 40
Silicon 80 2.4
Soap 60 3,840.0
Soybean 50 40.0
Stearic acid 25
Sugar 30
Sulfur 15 1.6
Thorium 5 0.004
Titanium 10 0.008
Uranium 45 0.004
(Continued)
Static Electrification 117

TABLE 7.3 (Continued)


MIE for Some Dust Clouds and Layers
Dust Cloud Dust Layer
Material (mJ) (mJ)
Urea resin 80
Vanadium 60 8.0
Vinyl resin 10
Wheat flour 50
Wood flour 20
Zinc 100 400.0
Zirconium 5 0.0004

Source: Data from the US Bureau of Mines. Reprinted with permis-


sion from IEEE Std. 142-2007.

TABLE 7.4
The Tribo-Electric Series
Positive
1 Air 13 Paper 25 Acrylic
2 Human skin 14 Cotton 26 Polyester
3 Asbestos 15 Wood 27 Celluloid
4 Glass 16 Steel 28 Orlon
5 Mica 17 Sealing wax 29 Polyurethane
6 Human hair 18 Hard rubber 30 Polyethylene
7 Nylon 19 Mylar 31 Polypropylene
8 Wool 20 Epoxy 32 PVC
9 Fur 21 Copper, nickel 33 Silicon
10 Lead 22 Silver, brass 34 Teflon
11 Silk 23 Gold, platinum Negative
12 Aluminum 24 Spongy polystyrene

appear when two similar materials are separated as often happens when plastic bags
are opened or a sheet of “cling film” is drawn from a roll. In such cases the surface
charging is attributed to the transfer of ions due to the presence of impurities.

7.3.1  Contact Electrification between Solid Surfaces


(A) Contact between two dissimilar metals
In this case free electrons move from the surface of the metal with the
lower work function1 to the one with the higher work function. As a result
1 The work function is the energy required to remove a free electron from the surface of a pure metal in
vacuum and is measured in electron-volt (eV).
118 An Introduction to Safety Grounding

(b)
φ1

φ2

φ1 > φ2
(c)
(a)

FIGURE 7.2 (a) Formation of surfaces charges due to differences in work function φ; (b)
surface roughness; (c) enhancement of electric field at sharp points.

of this transfer there will be, at equilibrium, a layer of negative charges on


one of the surfaces and a layer of equal positive charges on the other surface
as shown in Figure 7.2a.


In this case too, positive and negative charges are formed over the sur-
faces although not because of a difference in work function but because
of the transfer of electrons (possibly by tunneling) as well as of positive
and negative ions from one of the surfaces to the other. However, the exact
mechanism of transfer is not well defined [14].


The transfer of charge in this case depends on the basic structure of the
insulating materials and the type and amount of ionic contamination on the
surfaces and in the atmosphere, as well as on the densities of donors and
acceptor impurities present in each material.


When two charged surfaces are separated the capacitance between them
decreases, and this is accompanied by a large increase in the potential dif-
ference since the charge remains constant in the relationship

V = Q /C

The external work done to separate the surfaces appears as energy stored in the elec-
trostatic field (½QV). In practice, however, the separation of two surfaces is accom-
panied by a reduction in surface charge. There are two reasons for this:
The first reason is that on a microscopic scale surfaces are far from smooth (Figure
7.2b) so that during the separation process charges tend to accumulate at the last
remaining points of contact between the two surfaces. Because such points can be quite
sharp they produce a field of very high intensity in the gap (Figure 7.2c) leading to the
discharge of surface charges through such points by field emission, a process referred
to as field emission back discharge. The speed of this back discharge will depend upon
the speed with which the charges move to the discharge points and this, for a given
separation velocity, depends upon the surface resistivity. The lower this resistivity the
Static Electrification 119

smaller the charges left on the surfaces after separation. Also, the faster the separation
the greater the number of charges that will be left on the surfaces.
The second reason for the reduction of surface charges on separation is that during
the separation the field in the gap between the two surfaces (depending on the initial
surface charge density) may attain a value of 30 kV/cm, which is the breakdown
strength of air and is sufficient to produce ionization and initiate local discharges.
The positive and negative ions resulting from such a process will be drawn to the
surfaces and neutralize their opposite charges there.
Because of the above-mentioned reasons it is difficult to predict the amount of
charge that will remain on surfaces after their separation. Table 7.5 gives some typical
values for the electrostatic voltages which can be generated by tribo-electrification;
the magnitude of these voltages is strongly dependent on the relative humidity (RH),
but voltages of 10 and 20 kV are certainly possible.

7.3.2 Static electrification in liquidS and gaSeS


When a metallic surface comes in contact with an electrolyte which has a high relative
permittivity (such as water, εr = 80) an exchange of ions takes place between the two
media leading to the formation at the interface of a layer of positive charges and a layer
of negative charges known as the Helmholtz double layer as shown in Figure 7.3.
These charges become separated from each other when the liquid flows. For
example, if the liquid flows in a metal pipe and is collected in a tank which is insu-
lated from the pipe, charges accumulate in the tank and create an ignitable atmo-
sphere. An example of the latter is the washing of empty oil tanks of tanker ships;

TABLE 7.5
Some Typical Statically-Generated Voltages
Voltage(V)
Relative Humidity
Means of Static Generation 10%–20% 65%–90%
Walking across carpet 35,000 1,500
Walking on vinyl floor 12,000 250
Worker moving at bench 6,000 100
Opening a vinyl envelope 7,000 600
Picking up common polyethylene bag 20,000 1,200
Sitting on chair packed with polyurethane foam 18,000 1,500

metal liquid

FIGURE 7.3 Formation of Helmholz double layer at metal/liquid interface.


120 An Introduction to Safety Grounding

these tanks always have a flammable atmosphere of hydrocarbon gases/air mixture.


Spraying with statically charged water during washing creates a tribo-electrically
charged mist which under certain favorable conditions can cause extensive ESDs
within the tank which are known to have been the cause of explosions. This phenom-
enon, which is known as flow electrification, can occur with insulating liquids which
have a high content of moisture or other electrolytic impurities.
Static electrification in liquids can also occur at the interface between two liquids,
especially when this surface is very large as it is in the case of emulsions of oil (or
any other hydrocarbon liquid) and water.
In the case of clean gases flowing in metal pipes there is no static charge genera-
tion. However, if the gases contain entrained dust particles or aerosols then charges
may form.

7.4 WAYS FOR REDUCING THE FORMATION


OF SURFACE CHARGES
It has been mentioned in Section 7.3.1 that the charge which appears on the surfaces
of two metals during their contact depends essentially on the difference in their work
functions. It is therefore possible to reduce surface charges by choosing metals with
as small a difference as possible in their work functions. The work function of a
number of metals is given in Table 7.6.
In the case of insulating surfaces the materials may be chosen so that they lie
close to one another in the tribo-electric series, although this may not always be a
guarantee. Perhaps the best method for minimizing the charges which appear on
the separation of such surfaces is to reduce their surface resistivities. Experimental
investigations [10] have shown that for a large number of plastic materials and a nor-
mal separation speed of 1 m/s, surface charges could only be detected on surfaces
whose resistivities exceeded 1012 Ω/sq.2

TABLE 7.6
The Work Function (eV) of Pure Metals
Li 2.48 Be 3.32–3.92 Th. 3.38 Fe 4.49
Na 2.28 Mg 3.67 C 4.35–4.60 Ni 4.96
K 2.22 Ca 3.20–3.71 Si 3.54 Pd 4.98
Cs 1.93 Ba. 2.51 Ta 4.13 Pt 5.36
Cu 4.45 Zn 4.29 C 4.60
Ag. 4.46 Al 4.20 M 4.24
An. 4.89 Zr 3.73 W 4.54

2 The surface resistivity is the ratio of the dc voltage drop per unit distance between two parallel elec-
trodes in contact with the surface to the surface current per unit electrode length. Its unit is ohms, but to
avoid confusion with surface resistance, it is often expressed as ohm/square. It is numerically equal to
the surface resistance between opposite sides of a square of any size when the current flow is uniform.
Static Electrification 121

According to their surface resistivities materials are classified into three groups:

• > 1012 Ω/sq nondissipative


• 1010 –1012 Ω/sq antistatic (reduce the amount of charge generated by contact
and separation)
• 105–109 Ω/sq statically dissipative (antistatic)
• 100–105 Ω/sq conductive (no charge accumulation)

The surface resistivity of some copolymer plastics can be reduced to 100–1,000 Ω/


sq by the addition of fillers such as carbon powder, carbon fibers, or stainless steel
fibers. Applications of conductive plastics include their use in medical devices, for
packaging of electronic devices and for ESD protection, as conductive storage con-
tainers for hazardous liquids, and wherever protection from ESD is needed.
One family of copolymer plastics, e.g., polyanilines, have a low surface resistivity
in the range of 100–1,000 Ω/sq and are known as inherently conducting polymers.
They are used mainly in antistatic applications, batteries, and photochemical cells;
their main drawback, so far, has been their processability and thermal stability.
Table 7.7 gives typical values of the surface resistivities of a number of commonly
used plastic materials.
One of the most important factors which influence the surface resistivity of many
materials is the relative humidity of the surrounding atmosphere and the capability
of the material to absorb moisture. Surface resistivities should therefore be measured
for the same relative humidity values as those expected in actual use. Some poly-
mers, especially silicones, have a very low moisture absorbency and their surface
resistivity is unaffected by humidity.

TABLE 7.7
Surface Resistivity of Some Plastics
Material ρs(Ω/sq)
CA cellulose 1012–1014
FEP fluorinated ethylene propylene copolymer 1016
PA 6 polyamide—nylon 6 5 × 1010
PA 12 polyamide—nylon 12 1013
PC polycarbonate 1015
PC polycarbonate—conductive 100–500
PE polyethylene—carbon-filled 103–104
PE(HD) (polyethylene—high density) 1013
PE(LD) (polyethylene—low density) 1013
PMMA (polymethylmethacrylate) 1014
PP (polypropylene) 1013
PS (polystyrene—conductive) 102–107
XLPS (polystyrene—cross-linked) ˃1015
PTFE (polytetrafluoroethylene) 1017
PVC (polyvinylchloride) soft 1013
PVDF (polyvinylidenefluoride) 1013
122 An Introduction to Safety Grounding

7.5 ELECTROSTATIC INDUCTION


Figure 7.4 shows how induced charges appear on a conducting body. When an
isolated charged body carrying a negative charge for example is brought near the
conductor, a positive charge appears on the surface of the conductor nearest the
charged body and an equal negative charge will appear on the far surface since the
total charge on the conductor is zero. If we assume that the conducting body is now
momentarily connected to ground as shown in Figure 7.4a, the induced negative
charges will flow to earth but the positive charges will remain unchanged since
they are bound to the negative charges on the charged body (Figure 7.4b). If the iso-
lated charged body is now removed from the vicinity of the conductor, the positive
charges are now free to redistribute themselves over the surface (Figure 7.4c). We
thus see that a conducting body can acquire a charge without making direct contact
between it and any charged body.
When the conductor which has acquired an induced charge approaches another
conducting body the whole or part of this charge will flow to ground via an arc dis-
charge (Figure 7.4d). If the body is not grounded, the discharge current will flow to
earth through the capacitance between this body and ground.

7.6 TYPES OF ELECTRIC DISCHARGES


There are basically three types of electrical discharges which occur in air (or any
other gaseous medium):

charged conducting
body body

FIGURE 7.4 (a–c) Charging of a conducting body by induction; (d) ESD.


Static Electrification 123

• Spark or arc discharges


• Brush and propagating brush discharges
• Corona discharges

The first two are potentially dangerous since they can cause damage to electronic
equipment and explosions in combustible atmospheres, while corona discharges have
too low an energy to cause ignition and are in fact used as a preventive measure
against charge accumulation.

7.6.1 the Spark or arc diScharge


When the electric field in the gap between two conductors reaches the breakdown
strength of air (30 kV/cm at normal temperature and pressure) there is an electrical
discharge which completely bridges the gap in the form of a spark (which initiates an
arc if the electrodes are connected to a supply). The energy stored in the interelec-
trode capacitance (½CV2) is dissipated in the spark discharge. If the discharge occurs
in an explosive atmosphere ignition may result if the discharge energy is greater than
the MIE of the medium. Of major concern too is the damage which such discharges
can cause to sensitive electronic equipment. Sparks occur between charged conduc-
tors. Since human beings are electrically conducting they can carry large induced
charges and as such are the prime source of spark discharges.
As a practical example suppose that a person wearing rubber-soled shoes is walk-
ing on a nylon or woolen carpet. From Table 7.4 we deduce that the soles will acquire
a negative charge, and since the body is a conductor, a positive-induced charge will
appear on the soles of the person’s feet and a negative charge on the upper part of
his body, especially at the tips of his fingers (Figure 7.5). As his hand approaches a
grounded metallic body an ESD will take place between the fingers and the metallic
body when the electric field between them exceeds the breakdown strength of air
(30 kV/cm). Humans cannot feel a discharge of less than 3500 V and discharges at
potentials greater than 25 kV are painful.
Because the human body is a conductor it acts as a capacitor in which charge can
be stored. Now the absolute capacitance of any conductor is its capacitance as an

woolen carpet rubber-soled shoes

FIGURE 7.5 Discharge from person to metallic body.


124 An Introduction to Safety Grounding

isolated body, i.e., between the conductor and infinity. The simplest example is the
capacitance of an isolated conducting sphere of radius r:

C = 4πε or

C = 111rpF

Since in general the capacitance of a body is a function of its surface area, it is pos-
sible to consider the human body as equivalent to a spherical conductor of diameter 1
m; thus the absolute capacitance of the human body is approximately 50 pF. Typical
values are in the range 50–100 pF.
When other bodies are in the proximity of a conductor, there will be other capaci-
tances between the conductor and these other bodies; these are the mutual capaci-
tances. For the human body the mutual capacitances are those between the body and
surrounding walls and those between the soles of the feet and ground (Figure 7.6).
If all these capacitances are taken into consideration, the capacitance of the human
body will be in the range of 50 pF to 250 pF.
As mentioned in Chapter 1 the resistance of the human body can vary widely from
10 kΩ if the discharge occurs from the tip of a finger to 1,000 Ω if from the palm of
the hand, down to 100 Ω if via a large metal object held in the hand.
To simulate a discharge from the human body the circuit shown in Figure 7.7a is
used for test purposes:

Cb = capacitance of human body


Rb = resistance of human body
Vb = charging voltage

A primary requirement is that the circuit inductance should be less than 0.1 μH.
Different manufacturers may use different values for the above parameters, but the
following values are those specified by the IEC3 and by the US Military Standards4
and are the ones commonly used:

IEC MIL
Cb 150 pF 100 pF
Rb 150 Ω 1500 Ω
Vb 15,000 V 15,000 V
½CV 2 16.9 mJ 11.3 mJ

Figure 7.7b shows a typical waveform of the discharge current. Its shape is that of
an impulse wave with front-time values between 200 ps and 10 ns, and tail-time values
between 100 ns and 2 μs. The magnitude of the peak current depends on the discharge
voltage Vband may reach 40A if the discharge voltage is 20 kV. The front time of the

3 IEC 60801-2 (1991), EMC for Industrial Process Measurement and Control Equipment, Part 2:
Electrostatic Discharge Requirements.
4 MIL-HDBK-263B (1994), Electrostatic Discharge Control Handbook.
Static Electrification 125

FIGURE 7.6 Self and mutual capacitances of human body.

< 0.1 H

200ps – 10ns

100ns - 2μs

FIGURE 7.7 (a) ESD circuit for human body; (b) typical ESD current waveform.

wave and the discharge energy are the two most important parameters which deter-
mine the severity of the discharge. The extremely rapid rise of the wave front current
indicates that an ESD contains frequencies in the GHz range so that the inductance of
the ground circuit will be of primary importance. Discharges of only a few hundred
volts can cause damage to sensitive electronic circuits or cause them to malfunction.
126 An Introduction to Safety Grounding

To prevent an ESD from causing damage the discharge current must be prevented
from flowing through the circuit. The simplest way to accomplish this would be to
place the circuit inside a grounded metal enclosure which will divert the discharge
current to ground with all circuit grounds connected to the enclosure (Figure 7.8a).
In order to prevent the discharge from penetrating into the enclosure, the enclosure
should have no holes or apertures whatsoever. In practice this is not possible because
holes and apertures are needed for the entry/exit of power and signal cables as well
as for ventilation. In the presence of such openings the ESD discharge will create
intense electric and magnetic fields around them which may lead to secondary dis-
charges to the internal circuitry (Figure 7.8b). A case of particular interest is when
a cable with an external ground connection penetrates the enclosure (Figure 7.8c).
Since, as mentioned above, the ESD current contains frequencies in the GHz range
the connection to ground will be predominantly inductive. Because of this the poten-
tial of the enclosure will rise momentarily to thousands of volts, but since the circuit
has an external ground connection, its potential will remain zero. The large potential
difference which thus appears between the circuit and its enclosure can produce a
secondary arc which will traverse the circuit. This can be prevented by connecting
the circuit to the enclosure. A complete discussion of the corrective measures for the
prevention of ESD damage is beyond the scope of this book and interested readers
will find an excellent account of this subject in Ott’s classic work [17]. Suffice it to say

intense EM field at
apertures

circuit circuit

(a) (b)

external grounding
conductor

circuit

secondary
arc

(c)

FIGURE 7.8 ESD to enclosure: (a) no openings; (b) with openings;(c) with external ground
connection.
Static Electrification 127

here that with a grounded enclosure and good design ESD current can be prevented
from flowing through the electronic circuitry.

7.6.2 BruSh diSchargeS and propagating BruSh diSchargeS


Brush discharges are discharges which occur between charged insulating surfaces
and nearby grounded conductors such as equipment or personnel (Figure 7.9). They
take the form of filamentary discharge channels (streamers) which look like a brush
under low-light conditions (see also corona discharge in the next section). Since
the charges on the insulator surface cannot move a single discharge channel is not
formed. The maximum energy associated with such discharges is low and does not
exceed 4 mJ. However, this may be sufficient to ignite vapors of flammable liquids
which have MIEs less than 4 mJ (see Table 7.2).
A potentially more dangerous type of brush discharge is known as a propagating
brush discharge. It occurs when a grounded metallic surface has a highly charged
insulating layer in the form of a thin film or a deposit of dust or powder (Figure 7.10a).
The combination forms a capacitor5 that can store a large amount of energy; when
the potential difference across the insulating layer exceeds its breakdown strength,
it is punctured. The field parallel to the surface then becomes sufficiently great for a
massive surface discharge to occur. Such discharges are also known as Lichtenberg
discharges and they are potentially very dangerous since their energy is of the order
of several joules enough to ignite flammable gases and vapors as well as combustible
dusts. Such discharges occur in vessels and pipes which have an insulating lining and
in metal surfaces exposed to dusts and powders. Experiment has shown that this type
of discharge does not occur for film thicknesses greater than 10 mm or if the break-
down voltage is less than 4 kV for films and less than 6 kV for fabrics.
In some cases a propagating brush discharge can occur over the surface of insu-
lators in the absence of a grounded metal surface (Figure 7.10b). When charges are
formed on the inner side of a surface, as for example when a plastic bag is filled with
a charged powder or when a liquid flows inside a plastic pipe, these charges may
create an external field of sufficient strength either to attract charges of the opposite

filamentary streamers
– – – – –

charged insulating
surface
grounded conductor

FIGURE 7.9 Brush discharge.

5 Although a capacitor is usually associated with two conductors, a flat surface charge of uniform density
is considered the equivalent of a conducting plane surface.
128 An Introduction to Safety Grounding

electric puncture

insulating film 7
metal surface

(a)

bipolar charge
on thin film 7

(b)

film puncture site

surface discharge pattern


(c)

FIGURE 7.10 Propagating brush discharge: (a) insulating film on metal surface; (b) double
layer on insulating film; (c) Lichtenberg surface discharge pattern.

polarity or to ionize the air in the immediate vicinity of the surface. In either case a
double layer of charge will form across the insulating wall. If the insulation is punc-
tured a propagating brush-type discharge can occur.

7.6.3 corona diSchargeS


Corona is a phenomenon associated with highly nonuniform electric fields and is
characterized by the onset of detectable ionization and excitation long before break-
down occurs. Figure 7.11 shows the distribution of the electric field between a
grounded needle-like electrode and a charged surface. The field intensity around the
sharp tip is very much higher than that in the rest of the gap. When this field reaches
about 30 kV/cm it will produce a local discharge which manifests itself as a luminous
glow around the point with the microammeter indicating a flow of current. This glow
is known as corona. Briefly the mechanism is as follows.
Due to ultraviolet radiation, cosmic radiation, etc., a number of free electrons and
ions are always present in the atmosphere. When acted upon by an electric field they
will be accelerated and after traveling a short distance will collide with air molecules.
When the field is sufficiently high the result of such collisions is either the removal of
an electron from its molecule (ionization), or the raising of an electron into a higher
orbit (excitation). Such orbits are unstable and the electron will drop back to its origi-
nal orbit and in so doing release the energy originally acquired; this energy appears
Static Electrification 129

corona region

+
+
+
+
A

FIGURE 7.11 Electric field between a sharp point and a charged plane.

as visible light or corona. The flow of current is the result of the movement of the
generated electrons and ions toward the oppositely charged electrodes. Coronas can
be “noisy” both acoustically and electromagnetically.
It should be stressed here that corona is invariably associated with highly non-
uniform fields and the energy dissipated is always too low to initiate an explo-
sion. In nonuniform fields where electrodes are not sharp the corona discharge is
replaced by a brush discharge which can be considered as a type of high-energy
corona discharge.
Since corona is a field effect it can occur either around a live electrode or around
a grounded electrode. To distinguish between these two cases it is usual to refer to
coronas around live electrodes as active coronas and those around grounded elec-
trodes as passive coronas.
Besides their use in a large number of industrial applications, corona discharges
are used as an effective means of discharging statically electrified surfaces.

7.7 METHODS OF CONTROLLING ELECTROSTATIC CHARGES


7.7.1 grounding and Bonding
The accumulation of charges on conducting bodies can be effectively controlled by
grounding. All parts of an equipment or device should be bonded together and the
entire system connected to the supply ground. Grounding allows charges to leek to
ground but at the same time if a ground fault occurs during the operation of equip-
ment the potential of any grounded parts will rise to If × Rg above true ground. When
all parts are connected together (bonded) and grounded, they will all be at the same
potential and there is no danger of a discharge occurring either between them or to
ground.
As aforementioned the human body is a good conductor capable of acquiring
a charge sufficient to cause a spark discharge which has the energy to ignite flam-
mable material and damage electronic components. As a preventive measure person-
nel working in the manufacture and assembly of sensitive electronic components and
equipment, or other static-sensitive industries, are provided with wrist strap made of
conductive material and connected to ground through a 1 MΩ resistor; this resistor
130 An Introduction to Safety Grounding

will limit the current to about 0.25 mA should the wearer accidentally come in con-
tact with 230 V. Such a wrist strap is effective for the dissipation of static charge
acquired by the human body but not charges on any items of clothing. Where the use
of wrist straps is not practical because of worker mobility requirements, static control
flooring is used.

7.7.2 floorS and clothing for Static control


These are floors made from a variety of materials which are made conducting by the
addition of carbon in some form (graphite, fibers) or special chemical compounds
and include polymers, epoxy based mortars, hard rubbers, etc., either as a continuous
layer or in the form of tiles. Their function is to dissipate static charges by ground-
ing all equipment and personnel in direct contact with the floor. They will also sig-
nificantly reduce the body voltage generated by tribo-electrification. Such floors are
used wherever the occurrence of ESDs can have serious consequences such as in
the explosives industry, hospital operating rooms, aircraft hangers, manufacture and
assembly of sensitive electronic units, or in any areas where there are flammable
vapors or gases susceptible to static ignition. There are two types of static control
flooring classified according to their resistance to ground: static conductive floors
and static dissipative floors. The distinction between the two types is based on the
resistance to the movement of charge across the material’s surface. Conducting floors
have a much lower resistance than dissipative floors. For the resistance values to be
significant they must always be referred to the standard method of measurement
used; here we give (Figure 7.12) the most common and practical method as specified
by the Standards ASTM F150-06 and ANSI/ESD S7.1-2013.

In both the above tests all electrodes should be a minimum of 90 cm away from any
grounded items or earth ground, and readings taken after 15 seconds. Resistance
values are the average of at least five different measurements. If the floor consists
of tiles then some of the PTP measurements must be made between electrodes
across the seam. Both tests can be carried out on individual floor specimens or on
installed floors. For the exact testing procedure the relevant standards should be
consulted.
Based on the above measurements the classification of floorings according to their
resistance is as follows.
Static Electrification 131

megger
2.5” (6.35 cm) 5 lb (2.27 kg)
diameter per electrode

3’ (91 cm)
conducting floor

seam if tiled floor

(a)

grounding point

(b)

FIGURE 7.12 Measurement of surface resistance: (a) PTP; (b) point-to-ground.

7.7.2.1 Static Conductive Floors

PTP resistance: > 1 × 104 Ω


Point-to-ground resistance: < 1 × 106 Ω

These limits have been chosen because if the floor is too conductive personnel may
suffer electric shock if exposed to mains voltage, whereas if it is insufficiently con-
ductive charge may not be removed completely or rapidly enough.

7.7.2.2 Static Dissipative Floors

Point-to-ground resistance: > 1 × 106 Ω and < 1 × 109 Ω

Dissipative floors are used wherever rapid rate of charge dissipation can create a
magnetic field which could pose problems with the manufacture of electronic com-
ponents. Otherwise conductive floors are the superior choice.
To be effectively grounded personnel have to wear nonsparking conductive foot-
wear known as ESD footwear. The static control system consists of the person, the
ESD footwear, the conducting floor and the connections to ground. The system resis-
tance, which is the resistance from a person’s hand to ground, is the sum of the indi-
vidual resistances Rbody + Rfootwear + Rfloor + Rground. According to the Standard ANSI/
132 An Introduction to Safety Grounding

ESD S20.20(2014),6 the wrist strap resistance to ground should not exceed 3.5 ×
107 Ω, the system resistance should be < 1 × 109, and a walking test is required to
ensure that the maximum voltage generated does not exceed 100V. The IEC 61340-
5-1(2007)7 requirements are quite similar: either,

• The total resistance from persons to equipment ground via footwear and
floor shall be less than 3.5 × 107Ω;
or,

• The maximum body generation (also called walking body voltage) shall be
less than 100 V and the total resistance of the system shall be less than 1 ×
109 Ω.

It is worth mentioning here that it has been found impossible for a floor with a system
resistance < 35 MΩ to generate more than 100 V of static electricity.
It should also be pointed out here that in some applications where hazardous
atmospheres exist, such as anesthetizing locations and the handling of explosives,
the system resistance has to be less than 106 Ω. For example, DOD 4145.26M (2008)8
specifies that Rsystem ≤ 106 Ω and that the resistance from floor (or table top) to ground
must be > 40 kΩ for 110 V supply and >75 kΩ for 220 V supply.
In addition to personnel, mobile equipment, and items of furniture such as chairs
and tables should be of conducting material and directly grounded or grounded
through conductive rubber castors.
Since the resistance of static control floors and ESD footwear varies with time
and with use, their resistances must be periodically measured to ensure that they
still comply with their initial specifications and a record should be kept of such
measurements.
Figure 7.13 shows the essential features of a typical workstation. They include a
static dissipative work surface and ground mat, a personnel grounding wrist strap, a
common grounding connection, and appropriate labeling.
In addition to footwear due consideration has to be given to the type of cloth-
ing worn by personnel. Electric charges are generated on operators’ clothing by
tribo-electric effects (rubbing against upholstery, chairs, etc.) but because usual
clothing is electrically isolated from the body these charges cannot be dissipated
to ground via the skin. To overcome this problem special ESD protective clothing
is commercially available. Nonstatic fabric composition varies and is typically
as follows:

Polyester 30%–80%
Cotton 30%–75%
Carbon fiber 1%–5%

6 ANSI/ESD S20.20, Development of an ESD Control Program


7 IEC 61340-5-1, Protection of electronic devices from electrostatic phenomena—General requirements.
8 DOD 4145.26 M, 2008, Safety Manual for Ammunition and Explosives, C6.4: Static Electricity and
Grounding.
Static Electrification 133

dissipative table top

wrist strap
1MΩ to ground

dissipative
footwear

dissipative floor
connected to earth
bonding point

FIGURE 7.13 Schematic diagram of a typical ESD workstation.

Whether special clothes are needed or not depends on the nature of the work envi-
ronment. In electrostatic protected areas, especially in clean rooms and very dry
environments, and where electronic parts being manufactured or handled have a very
high degree of sensitivity, ESD protective clothing is required. Such clothing is also
necessary in regions where the MIE of flammable vapors is less than 0.2 mJ.
ESD protective fabrics should have a surface resistivity of less than 5 × 1010 ohms.
The most quoted standard for the measurement of this resistivity is the European
Standard:
EN 1149-1: 2006, Protective clothing—Electrostatic Properties—Part 1: Test methods
for measurement of surface resistivity.

The US Standard for protective clothing is,


ANSI/ESD STM 2.1-2018: Electrostatic Discharge Association standard test method
for the protection of ESD susceptible items—Garments.

According to this test method the PTP resistance should be as follows:

Garments Rp-p< 1 × 1012 Ω


Groundable garments9 Rp-p< 1 × 109 Ω

7.7.3 control of humidity


As mentioned in Section 7.4 the RH of the ambient air has a marked effect on the
surface resistivity of materials and hence on the ability of the electrostatically gen-
erated surface charges to leak away. The higher the RH the easier it is to dissipate
these charges. As shown in Table 7.5 there is a dramatic reduction in body voltage

9 According to IEC 61340-5-1 “If a groundable garment is used as part of the person’s primary ground
path (person is connected to a garment which is connected to a grounding cord that is attached to
ground) then the maximum resistance from the person’s body to ground should be 3.5 ×107Ω”.
134 An Introduction to Safety Grounding

generated by tribo-electrification with increase in the RH. If raising the RH does


not adversely affect the manufacturing process then this is one of the most effective
ways of limiting static charging. The humidity may either be raised locally or in the
atmosphere as a whole. A RH of 60% – 70% is sufficient to minimize the dangers
associated with surface charging.

7.7.4 ionization
Air ionization is a very effective way of eliminating static charges on nonconductive
materials and isolated conductors and is extensively used for the control of ESD in
industrial (especially textile and paper industries), electronic, and clean-room envi-
ronments. As outlined in Section 7.5 a corona discharge is a source of ions generated
by the ionization of air in the immediate vicinity of a highly stressed point (or wire)
electrode. The net charge produced has the same sign as the polarity of the point.
In a passive corona discharge the polarity of the point will be opposite to that of
the surface charge. Thus for a negatively charged surface positive ions will move
away from the positive point toward the surface and neutralize charges there; con-
versely, if the static charge is positive negative ions will be generated at the point and
drawn to the surface. Passive corona will cease once the field at the point is reduced
below the value required to initiate corona, and there will therefore only be a partial
cancellation of surface charges. Passive ionizers usually consist of a grounded comb-
like metal structure consisting of a row of sharp points or teeth on a metal bar or can
have a brush-like structure with copper or bronze bristles resembling either a hair
brush or a bottle brush.
An active corona source is the best way to reduce static charges to a very low level.
The most common type of active ionization device consists of an AC high voltage (2–7
kV) transformer connected to a set of pointed emitters10 via decoupling resistor and
capacitor to ensure that currents drawn if the point is touched are not dangerous. Since
the voltage is alternating the corona discharge will generate ions of both signs thus
enabling the neutralization of either positively or negatively charged surfaces. However,
because such a discharge suffers from a high rate of recombination of positive and neg-
ative ions, discharge points should be located no more than 2–10 cm from the charged
surface. To reduce the probability of recombination at near distances from the point,
filtered compressed air or a fan is used to disperse the ions produced thus increasing the
effective range to as much as 70 cm. An active ionizer can neutralize ±1,000 V to ±100
V in less than 10 seconds. It can be switched on and off and can be located wherever
most needed. Many ESD workstations are provided with an active ionizer. However,
active ionizers must not be used in explosive environments unless specially designed to
guarantee that no sparking will occur in the event of a malfunction.
One disadvantage of active ionizers is that electrical discharges in air produce
ozone and nitric oxide which, if inhaled for long periods, produce chronic respiratory
irritation and therefore require proper venting.

10 Emitter materials are stainless steel, thoriated tungsten, or single crystals of silicon or germanium,
depending upon kind of application. Stainless steel has the highest erosion rate and single crystals
the lowest.
Static Electrification 135

Both passive and active ionizers are available commercially in diverse types and
sizes.
Some ionizers use a radioactive material, such as radium or polonium, which
emits alpha particles to ionize the air. However, the use of such ionizers is very
limited as they have to be licensed from the relevant government authorities and
strict safety measures have to be enforced. They can be safely used in explosive
environments and their size can be extremely small allowing their placement in very
confined areas. However, their half-life is relatively short (~130 days) so that they
have to be replaced periodically. Their use may also require periodic inspection by
the authorities.
8 Protection against
Lightning

8.1 NATURE OF LIGHTNING


The need to protect buildings and structures from the damaging effects of lightning
gave birth to protective grounding. The phenomenon of lightning itself has been
extensively studied and there are a large number of books and hundreds of research
papers devoted to it; however, it is relevant here to give the reader a brief outline of
the nature of lightning and the lightning stroke to enable him to better understand the
protective measures employed.
Lightning is generated by a special type of charged cloud known as a thundercloud.
A lightning discharge can occur inside one cloud, between two clouds, or between
a cloud and ground. The latter, known as a lightning stroke, is the most dangerous
type of discharge against which buildings and structures have to be protected.
Many theories have been put forward for the mechanism by which thunderclouds
are charged but we will confine ourselves to the results: the lower part of the cloud
carries negative charges and the upper part positive charges (Figure 8.1). As the
densities of these charges increase so does the electric field intensity associated
with them. When this intensity equals the breakdown strength of the ambient air a
discharge is initiated either inside the cloud, or between two clouds, or between the
cloud and ground.

discharge
between clouds

internal discharge

ground stroke

FIGURE 8.1 Types of lightning discharges.

137
138 An Introduction to Safety Grounding

Based on the evidence provided by high-speed photography of lightning strokes


we can describe the development of the stroke by the following sequence of ultra
rapid events:

1. The electric field between the cloud and ground reaches a value which
causes the formation of a negative streamer from the lower part of the cloud
directed toward the ground. This pilot streamer (or leader) is faint and diffi-
cult to see and since it follows the path of least resistance it appears stepped.
As it progresses downward it branches out into several stepped streamers
(Figure 8.2a).

3. When the upward and downward propagating streamers meet a large current
flows as a highly luminous discharge constituting the main or return stroke
(Figure 8.2c). This is followed by a series of subsequent discharges (four
strokes on the average) of much lower magnitude than the initial stroke and
resulting from the discharge of other portions of the cloud through the con-
ducting path created by the main stroke.

Since the distribution of positive and negative charges inside the cloud is not uniform,
the charge density will vary from one part to the other. Similarly the density of
induced charge over the surface of the ground will vary according to the conductiv-
ity of the soil (mineral ore, wet land, arid sand, etc.). Since a cloud base can cover an
area between 5 and 50 km2, the charge density over such an area will depend both on
the nature of the soil and on the ground topography. A lightning stroke will be more
likely to occur where the charge density is highest in the cloud and on the ground.
On the ground this will usually be some elevated point such as a tree, a transmission
line, a tall building, a mast, a spire, a person on an open field, etc.
Although the charges in a thundercloud cannot be controlled, it is however pos-
sible, by using an appropriate protection system, to direct the lightning stroke to

(a) (b) (c)

FIGURE 8.2 Steps involved in the development of a lightning stroke.


Protection against Lightning 139

line which defines


Ip slope of wave front
0.9Ip
current

0.5Ip

0.1Ip


time
Tf
Tt

Tf = time for current to reach its peak value (front time)


Tt = time for current to reach 50% of its peak value (tail time)

FIGURE 8.3 Waveform and parameters of an initial short lightning discharge current.

specific points on the ground. This is the function of all schemes for the protection
of structures against lightning.
Sometimes a positive streamer may start upward, e.g., from mountain peaks or
very tall buildings constituting a positive upward leader. However, downward flashes
represent the majority of lightning discharges.
Figure 8.3 shows the typical wave shape of an initial lightning stroke. The current
rises to a peak value (20–2,000 kA) within an extremely short time (1–20 μs) and
then decays with varying rates. Three parameters are needed to identify such waves:

Ip = peak value of current


Tf = time for current to rise to its peak value Ip (front time)
Tt = time for current to drop to 50% of Ip (tail time)

Thus a wave is defined as a Tf /Tt, Ip wave. Tf and Tt are measured from the virtual
origin Oʹ formed by the line joining the 0.1Ip and 0.9Ip points on the rising part of
the curve.
A 10/350 μs wave is used by both the IEC and IEEE standards to characterize the
current wave of the initial current stroke and a 8/20μs wave to characterize the cur-
rent wave of an indirect stroke.1

1 It should be mentioned that the overvoltages created by lightning strikes are characterized by a stan-
dard 1.2/50μs voltage wave which is used to test equipments’ withstand to overvoltages of atmospheric
origin.
140 An Introduction to Safety Grounding

8.2 WHY PROTECT AGAINST LIGHTNING


The average number of lightning strokes which hit the earth’s surface is 100 per
second. The energy of the discharge is dissipated as sound (thunder), heat, light, and
electromagnetic radiation.
When lightning strikes an unprotected building or a structure it causes very sub-
stantial damage. Metal parts are distorted by the huge electromechanical forces gen-
erated (proportional to the square of the current), concrete is shattered by heat, and
wooden structures are set on fire. The purpose of a lightning protection system (LPS)
is to draw the stroke to itself and provide a direct low-resistance discharge path to
ground.
What are the buildings and structures that have to be protected against lightning?
It is evident that structures which are liable to explode such as ammunition facto-
ries and depots, petroleum refineries, etc., have to be protected. As for other build-
ings and structures in order to evaluate whether protection is needed or not a risk
assessment has to be made. The type of risk philosophy and method of assessment
depends on the lightning protection standard to be used. The two most prominent
standards are IEC 62305-20102 and ANSI/NFPA 780-2017.3 Although a detailed
account of the procedure involved is beyond the scope of the present monograph we
have, however, found it pertinent to at least familiarize the reader with the processes
involved in risk assessment. In determining the risk factor The IEC Standard4 con-
siders that the types of loss that a structure and its contents can incur are of prime
importance:

• Human hazards, especially the risk of loss of life.


• Loss of production or service to public.
• Damage to structure and/or contents.
• Economic damage: physical damage to equipment and effect on insurance
premiums.

Having determined the type of loss a corresponding tolerable risk is determined from
tables. This risk is 10 –5/year for loss of life or permanent injuries and 10 –4/year for
other types of losses. A risk of 10 –5/year means a one in 100,000 chance of a light-
ning strike per year. The actual risk is then determined by a series of calculations
using formulae and weighting factors for several other aspects. If this actual risk is
less than the tolerable risk then no protection is needed.
The IEC standard defines four lightning protection levels (LPLs). For each level a
minimum current level to be protected against is designated together with the prob-
ability that the current may be greater than these levels:

2 IEC 62305-2010, Protection of Structures against Lightning Pt.1: General Principles; Pt.2: Risk
management; Pt.3: Physical damage and life hazards; Pt.4: Electrical and Electronic Systems;
Pt.5: Services.
3 NFPA 780-2017, Standard for the Installation of Lightning Protection Systems, Annex L: Lightning
Risk Assessment.
4 This standard has now been adopted by the UK as BS EN 62305 and has replaced BS 6651.
Protection against Lightning 141

Level Imin (kA) Probability Ip>Imin(%)


I 3 99
II 5 97
III 10 91
IV 16 84

I is the highest protection level and IV the lowest.


It is evident that the risk for a given structure will be a function of the frequency
of occurrence of lightning strikes at its geographic location and of the size (area) and
height of the structure. The higher the frequency and the larger and taller the building,
the greater the likelihood of its being struck. The geographical location determines the
number of thunderstorm days per year that occur in the region, a thunderstorm day
being defined as a day in which thunder is heard at least once. This number is known
as the keraunic (or ceraunic) level Td of the region. An isokeraunic map of the world is
shown in Figure 8.4 based on 1955 records of the World Meteorological Organization
(WMO); it is not necessarily 100% accurate today but it provides an example of such
maps and a useful guide if local meteorological records are not available.
The keraunic level Td is an indication of the thunderstorm activity (days per year
when thunder can be heard) but does not give an indication of the number of light-
ning strikes to ground which is of paramount importance when assessing the risk
to a structure. It has therefore been replaced by the flash density or lightning strike

FIGURE 8.4 Isokeraunic map of the world.


142 An Introduction to Safety Grounding

frequency Ng defined as the number of flashes to ground per year and per square
kilometer. When no information on Ng is available its value can be computed from
the relationship,5

N g = 0.04Td 1.25 (8.1)

To get the “feel” of risk assessment we give here the simple lightning risk assessment
as given in Annex L of the NFPA standard 780-2017.6
In addition to lightning flash density Ng, the other factors which must be taken into
consideration are the following:

• Structure location (environment)


• Type of construction
• Structure contents
• Structure occupancy
• Lightning consequences

The expected yearly strike frequency to a particular structure is given by

N D = N g × AD × C D × 10 −6 strikes/year (8.2)

where
Ng = flash density at structure location
AD = equivalent collective area of structure (m2)
CD = location factor
10 ‒6 is included because AD is in m2 whereas Ng is per km2.

The location factor CD takes into consideration the topography of the site; its value
for a structure of height H is given in Table 8.1 for various locations.

TABLE 8.1
Environmental Coefficient CD
Location CD
Structure surrounded by higher structures or trees within a distance of 3H 0.25
Structure surrounded by lower structures within a distance of 3H 0.5
Isolated structure, no other structures located within a distance of 3H 1
Isolated structure on a hilltop 2

5 R.R.Anderson and A.J.Eriksson, Lightning parameters for engineering applications,


Electra,69,(65-102),198.
6 Reprinted with permission from NFPA 780-2017, Standard for the Installation of Lightning Protection

Systems, Copyright © 2016, National Fire Protection Association, Quincy, MA. This reprinted mate-
rial is not the complete and official position of the NFPA on the referenced subject, which is repre-
sented only by the standard in its entirety which may be obtained through the NFPA website www.
nfpa.org.
Protection against Lightning 143

The effective collective area A D of a structure is the ground area having the same
yearly direct lightning flash probability as the structure. It is obtained by extending a
line of slope 1:3 (height of structure to horizontal collection distance) from top of the
structure to ground all around the structure as shown in Figures 8.5 and 8.6.
The tolerable lightning frequency NC is given by

( )
N c = 1 × 10 −3 / C events per year (8.3)

where C = C2 × C3 × C4 × C5.
The coefficients take into account the various factors given above and their values
are given in Tables 8.2–8.5. The quantity 1 × 10 ‒3 represents the acceptable risk factor
for property loss.
To determine if lightning protection is needed the tolerable lightning frequency
NC is compared with the expected strike frequency ND. If ND  >  NC protection is
definitely required; if ND ≤ NC protection is not needed (optional).

3H

H W
W
L 3H

1:3
H
3H L 3H

3H AD = LW + 6H(L + W) + 9πH
2

FIGURE 8.5 Equivalent collective area for a rectangular structure.

3H
H

Collective area 9πH 2

FIGURE 8.6 Equivalent collective area for a rectangular structure with a prominent part
which encompasses all portions of the lower part.
144 An Introduction to Safety Grounding

TABLE 8.2
Structural coefficient C2
Structure Metal Roof Nonmetallic Roof Combustible Roof
Metal 0.5 1.0 2.0
Nonmetallic 1.0 1.0 2.5
Combustible 2.0 2.5 3.0

TABLE 8.3
Structure Contents Coefficient C3
Structure Contents C3
Low value and noncombustible 0.5
Standard value and noncombustible 1
High value moderate combustibility 2
Exceptional value, flammable liquids, computer or electronics 3
Exceptional value, irreplaceable cultural items 4

TABLE 8.4
Structure Occupancy Coefficient C4
Occupancy C4
Unoccupied 0.5
Normally occupied 1
Difficult to evacuate or risk of panic 3

TABLE 8.5
Lightning Consequence Coefficient C5
Lightning Frequency IndexValue
Continuity of facility services not required, no environmental impact 1
Continuity of facility services required, no environmental impact 5
Consequences to environment 10

8.3 LIGHTNING PROTECTION SYSTEM


An LPS consists of three basic parts:

• A system of terminations on the roof and other elevated locations. This may
be either one or more connected vertical rods know as air terminals, or a
mesh of horizontal conductors known as air termination networks.
Protection against Lightning 145

• A system of down conductors to connect the roof terminations to the


grounding system.
• The grounding electrodes.

8.3.1 the air terminalS


The principle on which air terminals operate is quite simple. As mentioned in 8.1 the
main lightning stroke occurs when the negative downward streamer or leader from
a cloud meets the positive upward streamer initiated at some point on the ground.
The direction of the down streamer is not affected by conditions on the ground until
it reaches a certain distance from it which varies between 10 and 100 m. At this dis-
tance, which is known as the striking distance, the positive streamer will initiate and
move upward from some point at which the field intensity is greatest.
The function of the air terminal or lightning rod (Figure 8.7) is to attract the
downward leader. The negative charges at the base of the cloud draw the positive
charges induced on the ground to the tip of the terminal. These charges create an
intense field around the tip sufficient to ionize the surrounding air (this is the same
process which gives rise to a corona discharge and discussed in Section 7.5.3). This
ionization provides a medium of low resistance which facilitates the development of
the positive streamer and its upward propagation from the terminal tip toward the
downward streamer and the completion of the main stage of the lightning stroke.
Although the electric field at the tip of sharp ends is higher than that at blunt ends,
it decreases more rapidly with increasing distance; for this reason rounded tips are
today preferred to the pointed ones.

air terminal
with either
pointed or
rounded end

welded joint to
lightning conductor

FIGURE 8.7 Examples of an air terminal [Furse].


146 An Introduction to Safety Grounding

air terminal
h

cone radius

FIGURE 8.8 Cone of protection.

The region protected by an air terminal is known as the cone of protection


(Figure 8.8). The majority of standard specifications agree that the protected region
is the volume of the cone whose base radius is equal to the terminal height, i.e., a 1:1
cone of protection. All structures which lie inside the cone volume are completely
protected against direct strokes. Air terminals in the form of masts are used to protect
small isolated buildings, substations, and boats (see also Section 8.5).
Air terminals are usually placed on the roofs of structures to provide a zone of
protection around a building. The rods are typically installed around the periphery
of flat roofs within 24˝(60 cm) of end ridges and (as per NFPA 90) at intervals not
exceeding 20 ft (6 m) for terminals along the roof perimeter and 50ft (15m) for inner
terminals. All terminals are interconnected and connected to a number of down con-
ductors. By bonding together all the ground points of the down conductors around
the building, a ground ring is formed which encircles the building and helps to equal-
ize the potential of the entire earth system. An example of the use of air terminals on
the flat roof of a building is shown in Figure 8.9.

8.3.2 meSh air termination netWorkS


To protect structures which have a large surface area, an air termination network
is used. It consists of a number of horizontal conductors which form a mesh on the
roof surface and a number of vertical down conductors which connect the horizontal
conductors to the system ground. The design of a roof network depends on the
specification used. For example IEC 62305-3 specifies four mesh sizes 5 × 5 m, 10 ×
10 m, 15 × 15 m, and 20 × 20 m in descending order of the four LPLs (see Section 8.2),
with corresponding down conductor spacing 10, 10, 15, and 20 m, respectively. For
buildings higher than 60m the mesh network should also cover the vulnerable areas
Protection against Lightning 147

45m 15m
max max
6m
max

ground loop

FIGURE 8.9 Air terminals on flat roof.

mesh network as
close as possible to
edges

mesh size

FIGURE 8.10 Example of a mesh air termination network (down conductors not shown).

of the outer walls. With structures which consist of sections having different heights
with more than one roof, termination networks of all roofs should be bonded together
via their down conductors. Also all metal projections such as masts, aerials, air
conditioning units, etc., have to be properly bonded to the air termination network.
Figure 8.10 shows an example of an air termination network.

8.3.3 doWn conductorS


These conductors provide a low impedance path from the air termination network to
the grounding electrode system. When choosing and installing these conductors the
following recommendations should be observed:
148 An Introduction to Safety Grounding

1. The conductors have to follow the shortest path between the air network and
ground.
2. The conductors should be symmetrically mounted on the external surface of
the structure walls, starting from the exposed corners and distributed uni-
formly around the perimeter. Their spacing shall comply with the standard
specification used and shall in no case be more than 20 m.
3. It is best to avoid any re-entrant loops; however, if there are such loops then
they should comply with applicable separation distance s (see Section 8.4)
for the lengths shown in Figure 8.11.
4. The number of down conductors depends on the perimeter of the external
edges of the roof, but at least two down conductors are required for any
structure. Some specifications require that each down conductor be
connected to an earth electrode, while others specify that if the structure
perimeter is greater than 76 m then a ground connection is required every
30.5 m of perimeter or a fraction thereof, e.g., four connections will be
required for a perimeter of 96 m.
5. Each down conductor has to be provided with a test clamp to be placed
150 cm above the ground surface.
6. The conductors between the test clamps and the ground electrodes should
be at least 30 cm below ground and protected against corrosion, e.g., by
using PVC (polyvinylchloride) or XLPE (cross-linked polyethylene) -
coated conductors. Moreover, both the IEC and the NFPA recommend that,
in order to limit touch potentials, the last 3m of bare conductor be insulated
with at least 3mm of XLPE.

r ≥ 20 cm

90 o

l1

l2 s l1
s
l2
l3

preferred path

(a) (b)

FIGURE 8.11 Down conductor paths and separation distance s for (a) loop and (b) overhang.
(a) Acceptable if l2 ≥ s, l = l1 + l2 + l3; (b) l = l1 + l2.
Protection against Lightning 149

7. If reinforced concrete bars of the building are to be used as down con-


ductors they must form a continuous framework within the concrete and
equipotential bonding and earthing must comply with IEC 62305-1 or
NFPA-780 specifications. Close cooperation between builders and electri-
cal consultants is absolutely necessary from the design stage and throughout
the construction stage. If the electrical continuity cannot be guaranteed then
separate down conductors must be used.

8.3.4 typeS of conductorS


Air termination conductors and down conductors can be copper, aluminum,
or galvanized steel. In the case of aluminum or galvanized steel they have to be
connected to a copper grounding conductor (aluminum is not allowed) via special
bimetallic connectors to prevent corrosion. Moreover aluminum is liable to corrosion
when it is in contact with Portland cement and mortar mixes so that special clamps
must be used to prevent the conductor from touching the surface. Figure 8.12 shows
a selection of different types of clamps used for fixing down conductors to walls and
to structure elements.
The normal cross-sectional area7 of lightning conductors is 50 mm2. Conductors
can be either flat (strip) or solid round or stranded. Until recently the UK favored strip
conductors (typically 20 × 2.5 mm), Europe solid round conductors, and the USA

distance clamps surface clamp

I-beam clamp re-bar clamp pipe clamp

FIGURE 8.12 Fixing clamps for down conductors [Furse].

7 The section is chosen more for mechanical robustness than for current carrying capacity. Although the
current is very large its duration is extremely short. Thus from Eq. (5.1) for a current of 100 kA and
duration 50 μs with bolted joints (worst case) the minimum cross-sectional area is 6 mm 2.
150 An Introduction to Safety Grounding

stranded conductors. Round conductors have the advantage of being easier to handle
and install than flat conductors.
It is possible to use screened insulated cables with a semi conductive outer
sheath as down conductors. With these cables there is no risk of side flashing (see
Section 8.4) and minimize lightning-induced transients and are thus especially useful
if the premises contain sensitive electronic equipment.

8.3.5 grounding
Proper grounding is a vital part of any lightning protection system. All aspects
of grounding have already been considered in the first six chapters of this book.
The following additional recommendations should be considered for lightning
grounds:

• The resistance to ground should preferably be less than 5 ohms (National


Standards have to be consulted for the specified value).
• Exothermic welding of joints is preferred
• The top of ground rods should preferably be at a depth of at least 30 cm
below ground using an insulated conductor for the connection between elec-
trode and down conductor. This reduces the voltage gradient around the
electrode (see Section 2.1.1).
• It must be assured that the step and touch potentials are within safe limits
(see Chapter 4).
• All metal work on or around a structure must be bonded to the lightning
protection system to avoid side-flashing or else comply with the separation
distance requirements (see Section 8.4).
• Most standards recommend that the lightning grounding system and the
protective grounding system be connected together.
• To guarantee the effectiveness and durability (around 30 years) of the
protective system, the choice of material, installation procedures, and
workmanship must be of the highest quality and must fully comply with
the standard specifications that have been adopted. General inspection is
recommended every 1–5 years.

8.4 INDUCTANCE, SIDE FLASHING, AND


SEPARATION DISTANCE
The voltage to ground that appears on a down conductor during the wave front of the
lightning current stroke is given by

v ( t ) = iR + L di / dt (8.4)

where R is the conductor resistance and L its inductance. For down conductors the
inductance is of the order of 1.5μH/m which is quite small. However, because of
the extremely high rate of rise of the current wave front, the inductive component of
Protection against Lightning 151

the voltage becomes predominant. For example, for a front time Tf = 10μs, a peak
current Ip = 80kA, and a 25m long conductor,

25 × 1.5 × 10 −6 × 80 × 10 3
VL = = 300 kV
10 × 10 −6
This extremely high voltage may lead to an electrical flashover or breakdown between
some point on the down conductor and a separately grounded metal body such as a
water tank, air-conditioning unit, or any other component of an electrical system.
Such flashovers are known as side flashes (Figure 8.13). In order not to increase the
inductance of a down conductor almost closed loops, such as shown in Figure 8.11,
should be avoided if possible.
The rapid rate of rise of current also generates an equally rapidly changing
magnetic field which will induce transient voltages and currents in any installation
loops and cause damage to sensitive equipment. The induced voltage depends on the
mutual inductance M between the down conductor and the loop and is given by

di
Vi = M (8.5)
dt
and as di/dt is almost constant during the wave front, a square voltage will be induced
in the loop. The mutual inductance between a down conductor and a rectangular loop
(Figure 8.14), which could be a metal window frame, is given by

µo b s + a
M= ln (8.6)
2π s

water tank

side
flash
down conductor

FIGURE 8.13 Side flashing.


152 An Introduction to Safety Grounding

down conductor

voltage induced
in open loop
Vi

current induced
b
c in closed loop

FIGURE 8.14 Current or voltage induced in conducting loops by current in down conductor.

Assuming s = 1m, a = 1m, b = 0.5m, we find that M = 6.93μH.


For a 10/350μs, 80kA wave, di/dt= 8 × 109 A/s, so that

Vi = 8 × 10 9 × 6.93 × 10 – 6 = 55.4 kV

If the loop is open this voltage may cause a flashover across the gap. To protect
against such lightning-induced voltages sensitive equipment must be provided with
surge protection devices.
No side flashes will occur if all conducting parts are equipotentially bonded to
the lightning grounding system. However for equipment which is not bonded, such
as some conductive installations inside or on the roof of buildings, a separation
distance between the conductive item and the down conductor is necessary in order
to avoid side flashes.
According to the IEC 62305-3 standard, the separation distance is given by

kc
s ≥ ki l (8.7)
km

where
• ki is an induction coefficient which takes into account current steepness and
mutual inductance as well as the LPL (see 8.2):
LPL I ki = 0.08
LPL II ki = 0.06
LPL III and IV ki = 0.04
• kc is a factor that depends on the number of down conductors and the current
sharing between them. For buildings for which their length or width is equal
to 4 times their height the following values can be assumed:
2 conductors kc = 0.66
3 or more kc = 0.44
Protection against Lightning 153

I
2r a

● ●
● ●
d F
● ● magnetic flux of
● ● lateral currents
b
I

FIGURE 8.15 Electrodynamic force on a bent conductor.

• km is a factor which depends on the dielectric strength of the insulation material


km = 1 for air
km = 0.5 for concrete or bricks
• l is either the total length along the air termination and down conductors
from the point where the separation distance is to be considered to the
nearest equipotential bonding point or the distance between two points on
a down conductor which are closest. For example, in Figure 8.11(a), l =
l1 + l2+ l3 and in Figure 8.11(b), l = l1 + l2 .
Assuming a worst case scenario for which

ki = 0.08, kc = 0.66, km = 1

then if l = 25m, the separation distance required would be s = 1.32m.


Another factor to be considered, especially when there are inevitable bends in the
down conductors, is the very large electrodynamic force generated between sections
of a conductor. This force, repulsive if currents flow in opposite directions and attrac-
tive if in the same direction, is proportional to I 2.
The force acting on the section ab of the conductor shown in Figure 8.15 and due
to the magnetic field of the currents in the lateral parts of the conductor is given by

µo I 2 d
F≅ ln newtons (8.8)
2π r

for d = 10cm, r = 0.5cm, I = 80 kA, F = 3,835 N = 385 kg-force. This very strong
force, which tends to straighten the wire, is greatly reduced by replacing the sharp
corners with circular bends of radius ≥20cm (8″) as shown in Figure 8.11(a).

8.5 ZONES OF PROTECTION


The most widely used method for determining the areas of a structure or group of
structures that need protection and hence most appropriate locations of air terminals
or air termination network is the so called Rolling Sphere Method (RSM). The method
is based on the concept of the striking distance. As mentioned in Section 8.3.1 this
is the distance which a positive streamer, initiated at some point on a structure on
154 An Introduction to Safety Grounding

protected zones unprotected zones

FIGURE 8.16 Path of rolling sphere for identifying surfaces which need protection.

unprotected
zone

sphere radius
100 m
protected
zones

FIGURE 8.17 Protected and unprotected zones for a tall building.

the ground, travels upward to join the descending streamer, and produce the return
stroke. The initiation and propagation of the positive streamer from some point
depend on the field intensity between that point and the negative charges at the tip of
the downward leader reaching a critical value. The more the number of charges at the
tip, the greater the distance at which the critical field is attained and hence the greater
the striking distance and the higher the discharge current.
The tip of the downward leader is considered to be located at the center of an
imaginary sphere whose radius is equal to the striking distance. This sphere is then
rolled over the building (Figure 8.16) and any part that can be touched by the sphere
will be susceptible to a lightning strike. Those parts of the building that cannot be
touched are considered to lie within a zone of protection.
When there is a risk that a lightning strike to the side of a tall structure may cause
damage (Figure 8.17) then an extension of the air termination network to those parts
should be considered.
American standard NFPA 780 specifies a rolling sphere radius of 150 ft (45.7 m)
for normal use but this radius is reduced to 100 ft (30 m) for structures containing
Protection against Lightning 155

flammable liquids and gases. The new IEC Standard 62305-1 has four different
sphere radii corresponding to four LPLs mentioned in Section 8.2. These are

Imin Sphere Radius


Level (kA) (m)
I  3 20
II  5 30
III 10 45
IV 16 60

For each level the minimum peak current value (kA) is used to determine the
sphere radius S using the formula

S = 10 I 0.65 (m) (8.9)

The RSM may also be used to determine the protection zones in a substation. The
sphere radius is determined from the following formula adopted by the IEEE8

S = 8 k I 0.65 (m) (8.10)

  k = 1.0 for strokes to wire and ground plane,


= 1.2 for lightning strokes to masts,
I = return stroke current in kA.

Figure 8.18 shows an example of such an application. From the maximum height
of the substation equipment, the size of the substation and a rolling sphere radius,
the height and number of masts, and their distance apart can be determined. Mast
heights range from 3 to 15m and their number from 1 to 4. For further details the
reader is referred to the IEEE standard.

protected zone

FIGURE 8.18  RSM for identifying zones of protection in a substation with protective masts.

8 IEEE Std 998-2017, Guide for Direct Lightning Stroke Shielding of Substations.
156 An Introduction to Safety Grounding

In the 1:1 cone mentioned in Section 8.3.1 the protective angle is 45°. Such a cone
can be safely applied to determine the zones of protection for small substations and
simple-shaped structures of height up to 10m for level I protection and 30m for level
IV protection. In Figure 8.19 a 45° triangle is superposed on a scaled profile draw-
ing of the structure such that the cone apex coincides with the tip of the air terminal
whose protective zone is to be determined. Figure 8.20 shows a diagram drawn to
scale for comparing the zones of protection provided by a 10m high mast using a 1:1
cone and a 20m radius rolling sphere corresponding to the highest protection level;
the two zones are almost identical. For taller structures there are significant differ-
ences between the two methods and cone protective angles as small as 23° are neces-
sary. It should be mentioned here that the standard IEC 62305-1 includes a detailed
and somewhat laborious protection angle method in which, for each LPL, the cone
angle is given as a function of the height of the air rod above the ground plane for
each protection level.
In general the RSM is the preferred method used for identifying the surfaces of
structure that need protection because it simpler, safer, and can be used for all types
of structure and for all levels of protection required.

air terminals
section outside cone of
protection. Extra
terminal needed structure inside cone
45 of protection.
45

FIGURE 8.19 Cone of protection of a structure (profile must be drawn to scale).

S = 20m S = 20m

1:1 cone

45o

10m mast

FIGURE 8.20 Comparison between protected zone of a 1:1 cone of protection and rolling
sphere for a 10m mast (diagram drawn to scale).
Protection against Lightning 157

8.6 TANKS AND STACKS


Steel tanks with fixed or floating metal roofs and containing flammable liquids or
gases will not need protection against lightning provided that (a) all metal parts are
at least 3/16˝ (4.76 mm) thick and (b) their base is in intimate contact with ground all
around their perimeter to a depth of ≥50 cm. It must be ensured that the resistance to
ground is less than 5 ohms or has the value specified by the standard adopted.
Tanks with nonmetallic roofs have to be protected with air terminals, masts, or
aerial ground wires as dictated by experience and as required either by NFPA 780 or
by any other relevant National Standard.
All joints, couplings, and pipe connections have to be electrically continuous and
all vapor or gas openings closed or flame proof.
Stacks (chimneys) for flue gases and dusts must be protected by air terminals at
the top of the stack and uniformly distributed around the perimeter with a maximum
separation of 2.5 m. The length of the terminals depends on the type of gas or dust
emitted. It ranges from 0.5 m for nonflammable emissions to 1.5 m in the case of
explosive gases or dusts. If forced draught is used then the length of the terminals
should not be less than 4.5 m. Moreover, if the flue gases are corrosive a 1.6 mm thick
lead coating on the air terminals is required. In all cases a minimum of two down
conductors are required.
The above information should be used as a preliminary guide. For a comprehen-
sive protection policy the complete designated standard which is to be used must be
considered in its entirety.

8.7 PROTECTION OF TRANSMISSION LINES


BY AERIAL GROUND WIRES
Because overhead power transmission lines supported by tall towers run for long
distances through mostly flat territory, they are perhaps one of the structures that are
most vulnerable to direct lightning strokes. Overhead ground conductors can provide
a degree of protection against such strokes. One or more ground conductor is placed
on top of power transmission towers (Figure 8.21) and extends over the entire length
of the line. This conductor acts as a shield which protects the line conductors and
towers from atmospheric electricity and especially from direct lightning strokes. In
this section we give the theoretical background to clarify this.
Measurements have confirmed9 that under fair weather conditions there is a
constant electric field which extends vertically downward from the space charges
present in the upper atmosphere to earth. If the strength of this field is Eo then the
absolute potential (assuming that the potential at the earth’s surface is zero) at a point
P at a height z above the earth’s surface (Figure 8.22) is

Vo = Eo z (8.11)

9 B.F.J.Schonland, Atmospheric Electricity, Methuen, London, 1953.


158 An Introduction to Safety Grounding

ground conductor ground conductors

(a) (b)

FIGURE 8.21 Ground conductors on typical towers: (a) 66 kV; (b) 220 kV (dimensions in m).

Eo

V = Eo z

z
V=0

FIGURE 8.22 Fair weather atmospheric field.

and the equipotential surfaces in this case are horizontal planes. If an isolated
conductor is placed at any point in space it will acquire the potential of this point
but no electric charge. When the conductor is connected to ground it will acquire an
induced charge (see Section 5.7) and the secondary field produced by these charges
will modify the primary uniform as follows.

8.7.1 Secondary field of ground conductor


Assume that the charge per unit length on the ground conductor is λ C/m and that
the conductor is at a height h above ground. For the potential to be zero at all points
on the ground plane it is necessary to assume the presence of an image conductor
Protection against Lightning 159

λ r1
h P
z

r2
h
-λ conductor image

FIGURE 8.23 To determine potential at P due to a line charge and its image.

at an equal distance h below ground with an equal and opposite charge ‒λ C/m.
The potential at any point P (Figure 8.23) due to these two line charges is

λ r
VPg = ln 2 (8.12)
2πε o r1

At the ground surface where z = 0, r1 = r 2, the potential is zero according to Eqs.


(8.11) and (8.12); it must also be zero at the surface of the ground conductor, r1 = r,
r 2 ~ 2h, z~h (h>>r). We thus have that

λ 2h
(Vo + VPg ) = Eo h + ln =0
2πε o r

λ Eh
=− o
2πε o 2h
ln
r

It is evident that the charge which appears on the conductor is negative and by substi-
tuting this value in Eq. (8.12) we obtain

ln(r2 / r1 )
VPg = −Eo h (8.13)
ln(2h / r )

The total potential at P is thus

ln(r2 / r1 )
VP = VPo + VPg = Eo z − Eo h (8.14)
ln(2h / r )

The relative change in the potential at the point P due to the presence of the ground
wire is

VPo − Vp VPg h ln(r2 / r1 )


η= =− = (8.15)
VPo VPo z ln(2 h/ r)
160 An Introduction to Safety Grounding

If z = h' and assuming that the distances between conductors are small compared
with h we may write

r2 ≅ h + h′; r1 ≅ h − h′

and Eq.(8.15) becomes

 h + h′ 
ln 
h  r1 
η= (8.16)
h′ ln 2h
r
The ratio η indicates the effectiveness of the ground conductor in shielding the power
conductors from the atmospheric field. For example, if we assume the following prac-
tical values, r1 = 3 m, h′ = 19 m, h = 22 m, r = 6 mm.
We find that

22 ln(41 / 3)
η= = 34%
19 ln(44 / 0.006)
It should be pointed out here that the effect of the diameter of the ground conductor
and of its distance from a power conductor is limited since these distances appear in
the logarithmic term.
The electric field intensity in the direction of r1 is from Eq.(8.14)

∂VP h Eo
EP r 1 = − =−
∂r1 r1 ln(2h / r )

and on the surface of the conductor (r1 = r)

h Eo
Eg = − (8.17)
r ln(2h / r )

The negative sign indicates that the direction of the field is toward the conductor
surface; using the above values (h = 22 m, r = 6 mm) we find from the above equa-
tion that

3,667
Eg = Eo = 412Eo
8.9
The average value of the atmospheric field Eo under fair weather conditions is 100 V/m
and in the presence of thunderclouds it may rise to between 10 and 300 kV/m. If we
assume that Eo = 10 kV/m we find that

Eg = 4,120 kV/m = 41.2 kV/cm

This field is sufficient to produce a corona discharge around the conductor and draw
lightning to it thus protecting the power line.
Protection against Lightning 161

ground conductor
S
power conductor
h
y x

D2

D1

FIGURE 8.24 Ground conductor height and striking distance.

8.7.2 height of ground conductor


To determine the minimum height of the ground conductor we shall make use of the
RSM in which the radius of the sphere represents the striking distance S as described
in Section 8.5. If h is the height of the ground wire, y is the height of the highest line
conductor, and x is the distance between them, then from Figure 8.24 we have that

D12 = S 2 − ( S − h)2

D1 = h(2S − h)

Similarly,

D2 = y(2S − y)

so that

x = D1 − D2 = h(2 S − h) − y(2 S − y) (8.18)

For given values of S, y, and x, the minimum value of h can be determined from Eq.
(8.18). The striking distance can be determined from Eq. (8.10). As an example, for
I = 50 kA we find that

S = 8 × 50 0.65 = 100 m

and for y = 20 m, x = 4 m, S = 100 m, we find from Eq.(8.18) that h = 23 m.

8.8 PROTECTION AGAINST SURGES


In this chapter we have been mainly concerned with the protection of structures
against direct lightning strikes, a topic intimately associated with grounding which
is the main topic of this book. However, we must draw the attention of the reader
162 An Introduction to Safety Grounding

that there is a secondary effect of lightning which can cause extensive damage to
all types of equipment, particularly the sensitive electronic systems and devices
which permeate every aspect of life today. This secondary effect consists of induced
transient voltages and current surges which enter buildings through mains power
supplies, telephone and data communication lines, both underground and overhead.
Protection against such surges is today considered as an integral part of a lightning
protection scheme. Part 4 of IEC 62305 deals specifically with this aspect which lies
beyond the scope and subject matter of this book.
Appendix A: Wire Sizes

TABLE A.1
Cross-Sectional Area and Diameter of Copper Conductors (AWG)
Stranding Overall
Size
AWG/ Area Diam. Diam. Area
(kcmil) (cmil) Quan-tity (in.) (in.) (in.2)
18 1620 1 – 0.040 0.001
18 1620 7 0.015 0.046 0.002
16 2580 1 – 0.051 0.002
16 2580 7 0.019 0.058 0.003
14 4110 1 – 0.064 0.003
14 4110 7 0.024 0.073 0.004
12 6530 1 – 0.081 0.005
12 6530 7 0.030 0.092 0.006
10 10380 1 – 0.102 0.008
10 10360 7 0.038 0.116 0.011
8 16510 1 – 0.126 0.013
8 16510 7 0.049 0.146 0.017
6 26240 7 0.061 0.164 0.027
4 41740 7 0.077 0.232 0.042
3 52620 7 0.087 0.260 0.053
2 66360 7 0.097 0.292 0.067
1 83690 19 0.066 0.332 0.087
1/ 0 105600 19 0.074 0.373 0.109
2/0 133100 19 0.084 0.419 0.138
3/0 167800 19 0.094 0.470 0.173
4/0 211600 19 0.106 0.528 0.219
250 – 37 0.082 0.575 0.260
300 – 37 0.090 0.630 0.312
350 – 37 0.097 0.661 0.364
400 – 37 0.104 0.728 0.416
500 – 37 0.116 0.813 0.519
600 – 61 0.099 0.893 0.626
700 – 61 0.107 0.964 0.730
750 – 61 0.111 0.998 0.782
800 – 61 0.114 1.03 0.834
900 – 61 0.122 1.094 0.940
1000 – 61 0.128 1.152 1.042
1250 – 91 0.117 1.289 1.305
1500 – 91 0.128 1.412 1.566
1750 – 127 0.117 1.526 1.829
2000 – 127 0.126 1.632 2.092

163
164 Appendix A: Wire Sizes

TABLE A.2
Cross-Sectional Area of Copper Conductors According to German,a
British,b and USAc Standards
Nominal
Nominal Cross-Sectional Area Diameter Conductor Size
(mm2) (mm) Designation
0.65 1.1 NBS 3/.020
0.65 0.92 NBS 1/.336
0.75 1.02 0.75
0.821 1.12 AWG 18
0.97 1.2 NBS 1/.044
1 1.15 1
1.039 1.58 AWG 17
1.29 1.29 NBS 3/.029
1.307 1.4 AWG 16
1.5 1.65 1.5
1.652 1.99 AWG 15
1.94 1.63 NBS 3/.036
1.94 1.85 NBS 1/.064
2.083 1.8 AWG 14
2.5 2.08 2.5
2.625 2.21 AWG 13
2.9 2.34 NBS 7/.029
3.309 2.3 AWG 12
4 2.61 4
4.17 2.75 AWG 11
4.52 2.95 NBS 7/.036
5.26 2.8 AWG 10
6 3.3 6
6.45 3.71 NBS 7/.044
6.634 3.97 AWG 9
8.366 3.6 AWG 8
9.35 4.17 NBS 7/.052
10 4.67 10
10.55 4.88 AWG 7
13.296 5.2 AWG 6
14.52 5.23 NBS 7/.064
16 5.59 16
16.767 5.89 AWG 5
19.35 6.5 NBS 19/.044
21.15 AWG 4
25 25
(Continued)
Appendix A: Wire Sizes 165

TABLE A.2 (Continued)


Cross-Sectional Area of Copper Conductors According to German,a
British,b and USAc Standards

Nominal
Nominal Cross-Sectional Area Diameter Conductor Size
(mm2) (mm) Designation
25.81 6.61 NBS 19/.052
26.662 6.6 AWG 3
33.625 7.42 AWG 2
35 7.8 35
38.71 8.13 NBS 19/.064
42.406 8.43 AWG 1
48.5 9.3 NBS 19/.072
50 9.47 50
53.508 10.6 AWG 1/10
64.52 10.62 NBS 19/.083
67.442 11 AWG 2/0
70 11.94 70
77.5 12.8 NBS 37/.064
85.024 12.8 AWG 3/0
95 13.41 95
96.77 14.5 NBS 37/.072
107.218 14.61 AWG 4/0
120 14.8 120
126.675 16.2 MCM 250
129.03 16 NBS 37/.083
150 17.3 150
152.01 18 MCM 300
162 18.4 NBS 37/.093
177.345 18.49 MCM 350
185 19.56 185
193.55 20.5 NBS 37/.103
202.68 20.65 MCM 400
228.02 21.3 MCM 450
240 21.72 240
253.35 22.68 MCM 500
258.06 23.6 NBS 61/.093
278.71 23.6 MCM 550
300 24.49 300
304 25.35 MCM 600
322.58 NBS 61/.103
329.35 MCM 650
354.71 MCM 700
380 MCM 750
400 400

a VEB.
b NBS (New British Standard).
c AWG& MC.
166 Appendix A: Wire Sizes

TABLE A.3
Resistance of Copper Conductors
Cross-Sectional Area Resistance (70oC)
(mm2) (Ω/km)
0.75 29
1 21.7
1.5 14.7
2.5 8.71
4 5.45
6 3.62
10 2.16
16 1.36
25 0.863
35 0.627
50 0.463
70 0,321
95 0.232
120 0.184
150 0.15
185 0.1202
240 0.0922
300 0.0745
Bibliography
[1] G.T. Tagg, Earth Resistances, Newnes, London, 1964.
[2] V.Manoilov, Fundamentals of Electrical Safety, MIR Publishers, Moscow, 1975.
[3] W.F.Cooper, Electrical Safety Engineering, Butterworth, Oxford, 1994.
[4] Recommended Practice for Grounding of Industrial and Commercial Power Systems
(Green Book), IEEE Std. 142, 2007.
[5] The National Electrical Code, NFPA 70, 2017.
[6] Earthing, British Standard Code of Practice, CP 1013, 1965.
[7] Guide for Safety in AC Substation Grounding, IEEE Std 80-2000, (ANSI).
[8] Military Handbook, MIL – HDBK – 419A, 1987.
[9] A.P. Sakis Meliopoulos, Power Grounding and Transients, Marcel Dekker, New York,
1988.
[10] G. Luttgens and N. Wilson, Electrostatic Hazards, Butterworth, Oxford, 1997.
[11] L.B. Loeb, Static Electrification, Springer-Verlag, Berlin, 1958.
[12] Code of Practice for Control of Undesirable Static Electricity, BS 5958, 1991; Part 1:
General Considerations; Part 2: Recommendations for Particular Industrial Situations.
[13] Protective Clothing - Electrostatic Properties. Part 1: Surface Resistivity (Test Methods
and Requirements), EN 1149-1, 2006.
[14] Electrostatics Discharge Control Handbook, DOD – HNBK – 263, Dept. of Defense,
1980.
[15] Recommended Practice on Static Electricity, NFPA 77, 2019.
[16] IET Wiring Regulations, 18th Edition, BS 7671:2018.
[17] H.W. Ott, Noise Reduction Techniques in Electronic Systems, Wiley, New York, 1988.
[18] J. Cadick, M.C. Schellpfeffer and D. Neitzel, Electrical Safety Handbook, McGraw-
Hill, New York, 2000.
[19] Recommended Practice for Powering and Grounding Sensitive Electronic Equipment
(Emerald Book), IEEE Std. 1100, 2005.
[20] Standard for the Installation of Lightning Protection Systems, NFPA780, 2017.
[21] Guide for Measuring Earth Resistivity, Ground Impedance and Earth Surface
Potentials of a Grounding System, IEEE 81-2012.
[22] Low-Voltage Electrical Installations - Part 5: Selection and Erection of Electrical
Equipment - Earthing Arrangements and Protective Conductors, IEC 60364-5-54,2011.
[23] Protection against Lightning, British Standards BS EN 62305: 2010.
[24] Protection of Structures against Lightning, Parts 1–4, IEC 62305, 2010.
[25] Recommendations for the Protection of Structures against Lightning, W.J. Furse & Co.
Ltd., Nottingham, UK.
[26] Lightning Protection System Components, Parts 1–7, IEC 62561, 2017/2018.
[27] T.M. Kovacic, Electrical Safety, American Society of Safety Engineers, Des Plaines,
Illinois, 2001.
[28] R.P. O’Riley, Electrical Grounding, Delmar Publishers, New York, 1990.
[29] J.D. Cobine, Gaseous Conductors, Dover Publications, Inc, New York, 1958.
[30] R. Rudenberg, Transient Performance of Electric Power Systems, The M.I.T. Press,
1967.
[31] H.W. Denny, Grounding for the Control of EMI, Don White Consultants Inc., U.S.A.,
1983.

167
168 Bibliography

[32] J.A. Güemes and F.E. Hernando, Method for calculating the ground resistance of
grounding grids using FEM, IEEE Trans. Power Delivery, Vol. 19, No.2, pp. 595–600,
2004.
[33] T. Takashi and T. Kawase, Calculation of earth resistance for a deep driven rod in a
multi-layer earth structure, IEEE Trans. Power Delivery, Vol. 6, No. 2, pp. 608–614,
1991.
[34] F.P. Dawalibi, J. Ma and R.D. Southey, Behaviour of grounding systems in multilayer
soils: A parameter analysis, IEEE Trans. Power Delivery, Vol. 9, No. 1, pp. 334–342,
1994.
[35] C.L. Yaws, Matheson Gas Data Book, 7th Edition, Mcgraw-Hill, 2001.
[36] L.A. Lovachev, Flammability limits – a review, Combustion Science and Technology,
Vol. 20, pp. 209–224, 1979-Issue 5-6.
[37] V. Babrauskak, Ignition Handbook, Fire Science Publishers, Issaquah, WA, 2003.
Index
adiabatic equation 61 danger triangle 112
aerial ground wire 157ff depth of burial 9
air terminals 145 discharges, types of
air termination network 146 arc 123
anodic index 73 brush 127
apparent resistivity 20 corona 128
arc discharge 123 propagating brush 127
area of land 47 spark 123
atmospheric field 157 dissipative floors 131, 132
double insulation 91
bentonite 49 down conductor 147ff
bonding 129, 150 driven rod 23, 50
brazing 55 duration, current 4, 65
brush discharges 127 touch voltage 80
burial depth 9
buried plates 32 earth leakage breaker 83
buried wire 30ff electric shock 3
electrode
cable armor 60, 72 hemisphere 7
cable shield 60, 72 plate 32
cable trays 68 rod 23
capacitance, of human body 123, 124 wire 30ff
chemical grounding rod 49 electromagnetic induction 20
chemical salts 47 electrostatic charges 115
clamps 54, 149 electrostatic discharge (ESD) 115
clothing for esd 132 clothing 132
compression joint 56 damage 126
computer center 91 footwear 131
concrete 33 equipment grounding connections 47
encased electrodes 52 equipment grounding conductors 60
conductive floors 131 equipotential bonding 54
conductors equivalent hemisphere 24
down 146ff exothermic welding 55
equipment grounding 47 explosion limits 112ff
grounding 55, 60 extremely low voltage (ELV) 79
neutral 80ff
protective 80ff fall-of-potential method 11
cone of protection 146, 156 FELV 79
connections fence 76
mechanical 54, 149 fibrillation 4
welded 55 flash point 112, 113
contact electrification 111, 117 flexible metal piping 71ff
copperweld 50 floors, conducting 131
corona discharge 128 dissipative 131, 132
active 129, 134 flow electrification 120
passive 129, 134 footing resistance 40
corrosion 72 footwear (ESD) 131
current duration 4 force on conductor 153
physiological effects 3

169
170 Index

galvanic series 73 PELV 79


gas pipes 53 PME 87
ground connection provisions 60 power towers, grounding 40
ground grid 93 propagating brush discharge 127
ground mat resistance 94 protective grounding 79
ground potential rise (GPR) 93, 97 IT 89
grounding, at substations 74ff, 93ff PME 87
grounding conductor 56ff, 60ff TN-C 80
grounding systems 45, 79, 80 TN-S 84
TT 82
Helmholz double layer 119 protective earth conductor 81, 84, 87
hemispherical electrode 7 protective sleeve 59
hollow square 27
horizontal buried wire 30 reinforcement bars 33, 54, 149
relative humidity 133
ignition 111 resistance area 33
level resistance of human body 5
dusts 114, 116 resistance ratio 26
gases 113 resistance to ground 6, 45
liquids 113 resistivity
ignition sources 114 concrete 52
induced charge 122 improvement 47
inductance, mutual 151, 152 measurement 18ff
self 150 soil 16ff
ionization 134 surface 120, 121
ionizers 134 temperature dependence 18
isokeraunic map 141 risk assessment 142ff
isolation transformer 90 rod configurations
IT-system 89 circular 27
hollow square 27
jumpers 53, 72 multiple 25, 25, 28
single 23
keraunic level 141 solid square 27
rod length, effect of 24
lightning 137ff rods, resistance to ground
current wave 139 multiple 28
IEC protection levels 140 single 23
protection system 144 three in parallel 26
risk factor 142ff two in parallel 25
lower explosive limit (LEL) 112 rolling sphere method 153ff

Marconite 52 SELV 79
measurement of separation distance 150, 152
resistance to ground 11 side flashing 150
soil resistivity 18 soil resistivity 16, 17, 47
mesh sizes 146 soil treatment 47ff
mesh termination network 146 solid square 27
mesh voltage 98 spark discharge 123
minimum ignition energy (MIE) 112, 115 stacks 157
moisture, effect on resistivity 17 static conductive floors 130, 131
static dissipative floors 130, 131
neutral conductor 80, 87 step voltage 11, 35, 76, 98, 150
neutral grounding 74 Sverak formula 95
striking distance 145, 153
Onderdonk equation 57 substation grounding 74, 93ff
overhead ground wire 40, 157ff surface charges 115
oxygen 114 surface charging 115ff
Index 171

surface resistivity 120, 121 triboelectrification 111


swimming pool 79 TT-system 82

tanks 157 upper explosive limit (UEL) 112


thermit welding 55
TN-C 81 water pipes 33, 53
TN-S 84 welded connections 55
touch voltage 11, 36, 76, 98, 150 Wenner method 19
tower footing resistance 40 work function 117, 120
transfer voltage 39 wrist strap 129, 133
transformer substations 74ff
triboelectric series 117 zones of protection 153ff
Taylor & Francis eBooks
www.taylorfrancis.com

A single destination for eBooks from Taylor & Francis


with increased functionality and an improved user
experience to meet the needs of our customers.

90,000+ eBooks of award-winning academic content in


Humanities, Social Science, Science, Technology, Engineering,
and Medical written by a global network of editors and authors.

TAYLOR & F RANCI S EBOOKS OFFERS :

Improved
A streamlined A single point search and
experience for of discovery discovery of
our library for all of our content at both
customers eBook content book and
chapter level

REQUEST A FREE TRIAL


support@taylorfrancis.com

You might also like